Understanding Mechanical Ventilation: A Practical Handbook 2nd Edition

  • 65 3,818 4
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up

Understanding Mechanical Ventilation: A Practical Handbook 2nd Edition

Understanding Mechanical Ventilation Ashfaq Hasan Understanding Mechanical Ventilation A Practical Handbook Second Ed

5,771 916 8MB

Pages 554 Page size 198.48 x 320.64 pts Year 2010

Report DMCA / Copyright

DOWNLOAD FILE

Recommend Papers

File loading please wait...
Citation preview

Understanding Mechanical Ventilation

Ashfaq Hasan

Understanding Mechanical Ventilation A Practical Handbook Second Edition

Ashfaq Hasan 1 Maruthi Heights Road No. Banjara Hills Hyderabad-500034 Flat 1-E India [email protected]

ISBN: 978-1-84882-868-1

e-ISBN: 978-1-84882-869-8

DOI: 10.1007/978-1-84882-869-8 Springer Dordrecht Heidelberg London New York Library of Congress Control Number: 2010920240 © Springer-Verlag London Limited 2010 This work is subject to copyright. All rights are reserved, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilm or in any other way, and storage in data banks. Duplication of this publication or parts thereof is permitted only under the provisions of the German Copyright Law of September 9, 1965, in its current version, and permission for use must always be obtained from Springer. Violations are liable to prosecution under the German Copyright Law. The use of general descriptive names, registered names, trademarks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. Product liability: The publishers cannot guarantee the accuracy of any information about dosage and application contained in this book. In every individual case the user must check such information by consulting the relevant literature. Cover design: eStudio Calamar, Figueres/Berlin Printed on acid-free paper Springer is part of Springer Science+Business Media (www.springer.com)

‘To my parents’

Preface to the Second Edition

Simplify, simplify! Henry David Thoreau For writers of technical books, there can be no better piece of advice. Around the time of writing the first edition – about a decade ago – there were very few monographs on this subject: today, there are possibly no less than 20. Based on critical inputs, this edition stands thoroughly revamped. New chapters on ventilator waveforms, airway humidification, and aerosol therapy in the ICU now find a place. Novel software-based modes of ventilation have been included. Ventilator-associated pneumonia has been separated into a new chapter. Many new diagrams and algorithms have been added. As in the previous edition, considerable energy has been spent in presenting the material in a reader-friendly, conversational style. And as before, the book remains firmly rooted in physiology. My thanks are due to Madhu Reddy, Director of Universities Press – formerly a professional associate and now a friend, P. Sudhir, my tireless Pulmonary Function Lab technician who found the time to type the bits and pieces of this manuscript in between patients, A. Sobha for superbly organizing my time, Grant Weston and Cate Rogers at Springer, London, Balasaraswathi Jayakumar at Spi, India for her tremendous support, and to Dr. C. Eshwar Prasad, who, for his words of advice, I should have thanked years ago. vii

viii

Preface to the Second Edition

Above all, I thank my wife and daughters, for understanding. Hyderabad, India

Ashfaq Hasan

Preface to the First Edition

In spite of technological advancements, it is generally agreed upon that mechanical ventilation is as yet not an exact science: therefore, it must still be something of an art. The science behind the art of ventilation, however, has undergone a revolution of sorts, with major conceptual shifts having occurred in the last couple of decades. The care of patients with multiple life-threatening problems is nothing short of a monumental challenge and only an envied few are equal to it. Burgeoning information has deluged the generalist and placed increasing reliance on the specialist, sometimes with loss of focus in a clinical situation. Predictably, this has led to the evolution of a team approach, but, for the novice in critical care, beginning the journey at the confluence of the various streams of medicine makes for a tempestuous voyage. Compounding the problem is the fact that monographs on specialized areas such as mechanical ventilation are often hard to come by. The beginner has often to sail, as it were, “an uncharted sea,” going mostly by what he hears and sees around him. It is the intent of this book to familiarize not only physicians, but also nurses and respiratory technologists with the concepts that underlie mechanical ventilation. A conscious attempt has been made to stay in touch with medical physiology throughout this book, in order to specifically address the hows and whys of mechanical ventilation. At the same time, this book incorporates currently accepted strategies for the mechanical ventilation of patients with specific disorders; this should be of some value to specialists practicing in their respective ICUs. The graphs presented in this book are representative and are not drawn to scale. ix

x

Preface to the First Edition

This book began where the writing of another was suspended. What was intended to be a short chapter in a handbook of respiratory diseases outgrew its confines and expanded to the proportions of a book. No enterprise, however modest, can be successful without the support of friends and well wishers, who in this case are too numerous to mention individually. I thank my wife for her unflinching support and patience and my daughters for showing maturity and understanding beyond their years; in many respects, I have taken a long time to write this book. I also acknowledge Mr. Samuel Alfred for his excellent secretarial assistance and my colleagues, residents, and respiratory therapists for striving tirelessly, selflessly, and sometimes thanklessly to mitigate the suffering of others. Ashfaq Hasan, 2003

Contents

1

Historical Aspects of Mechanical Ventilation . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1 6

2

The Indications for Mechanical Ventilation . . . . . . . . 2.1 Hypoxia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 Hypoventilation . . . . . . . . . . . . . . . . . . . . . . . . . 2.3 Increased Work of Breathing . . . . . . . . . . . . . . 2.4 Other Indications . . . . . . . . . . . . . . . . . . . . . . . . 2.5 Criteria for Intubation and Ventilation . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

9 9 10 11 12 12 16

3

Physiological Considerations in the Mechanically Ventilated Patient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1 The Physiological Impact of the Endotracheal Tube. . . . . . . . . . . . . . . . . . 3.2 Positive Pressure Breathing. . . . . . . . . . . . . . . . 3.3 Lung Compliance . . . . . . . . . . . . . . . . . . . . . . . . 3.3.1 Static Compliance . . . . . . . . . . . . . . . . . 3.3.2 Dynamic Compliance . . . . . . . . . . . . . . 3.4 Airway Resistance . . . . . . . . . . . . . . . . . . . . . . . 3.5 Time Constants of the Lung . . . . . . . . . . . . . . . 3.6 Alveolar Ventilation and Dead-Space . . . . . . . 3.6.1 Anatomical Dead-Space . . . . . . . . . . . 3.6.2 Alveolar Dead-Space . . . . . . . . . . . . . . 3.6.3 Physiological Dead-Space . . . . . . . . . . 3.7 Mechanisms of Hypoxemia . . . . . . . . . . . . . . . . 3.7.1 Hypoventilation . . . . . . . . . . . . . . . . . . 3.7.2 V/Q Mismatch . . . . . . . . . . . . . . . . . . . .

19 19 21 28 29 32 34 38 39 40 40 40 46 46 50

xi

xii

Contents

3.7.3 Right to Left Shunt. . . . . . . . . . . . . . . . 3.7.4 Diffusion Defect . . . . . . . . . . . . . . . . . . 3.8 Hemodynamic Effects . . . . . . . . . . . . . . . . . . . . 3.9 Renal Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.10 Hepatobiliary and Gastrointestinal Effects. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.10.1 Hepatobiliary Dysfunction . . . . . . . . . 3.10.2 Gastrointestinal Dysfunction . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

52 54 55 60 62 62 63 63

The Conventional Modes of Mechanical Ventilation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 4.1 Mechanical Ventilators. . . . . . . . . . . . . . . . . . . . 71 4.1.1 Open-Loop and Closed-Loop Systems. . . . . . . . . . . . . . . . . . . . . . . . . . 72 4.1.2 Control Panel . . . . . . . . . . . . . . . . . . . . 72 4.1.3 Pneumatic Circuit . . . . . . . . . . . . . . . . . 73 4.1.4 The Expiratory Valve . . . . . . . . . . . . . . 73 4.1.5 Variables. . . . . . . . . . . . . . . . . . . . . . . . . 74 4.1.6 The Trigger Variable (“Triggering” of the Ventilator). . . . . . . . . . . . . . . . . . 75 4.1.7 Limit Variable . . . . . . . . . . . . . . . . . . . . 76 4.1.8 Cycle Variable . . . . . . . . . . . . . . . . . . . . 76 4.1.9 Baseline Variable . . . . . . . . . . . . . . . . . 78 4.1.10 Inspiratory Hold . . . . . . . . . . . . . . . . . . 79 4.1.11 Expiratory Hold and Expiratory Retard. . . . . . . . . . . . . . . . . . . . . . . . . . . 79 4.2 Volume-Targeted Modes . . . . . . . . . . . . . . . . . . 80 4.2.1 Volume Assist-Control Mode (ACMV, CMV) . . . . . . . . . . . . . . . . . . . 80 4.3 Intermittent Mandatory Ventilation. . . . . . . . . 84 4.4 Pressure–Support Ventilation . . . . . . . . . . . . . . 89 4.5 Continuous Positive Airway Pressure . . . . . . . 94 4.6 Bilevel Positive Airway Pressure . . . . . . . . . . . 97 4.7 Airway Pressure Release Ventilation (APRV) . . . . . . . . . . . . . . . . . . . . . . 97 4.7.1 Bi-PAP . . . . . . . . . . . . . . . . . . . . . . . . . . 98 4.8 Pressure-Controlled Ventilation . . . . . . . . . . . . 98 4.8.1 Proportional Assist Ventilation (PAV) . . . . . . . . . . . . . . . . . 101

Contents

4.9

Dual Breath Control. . . . . . . . . . . . . . . . . . . . . . 4.9.1 Intrabreath Control . . . . . . . . . . . . . . . 4.9.2 Interbreath (DCBB) Control . . . . . . . 4.9.3 Pressure Regulated Volume Control (PRVC) . . . . . . . . . . . . . . . . . . 4.9.4 Automode . . . . . . . . . . . . . . . . . . . . . . . 4.9.5 Mandatory Minute Ventilation (MMV). . . . . . . . . . . . . . . . 4.9.6 Volume Support (VS). . . . . . . . . . . . . . 4.9.7 Adaptive Support Ventilation (ASV). . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

Ventilator Settings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.1 Setting the Tidal Volume . . . . . . . . . . . . . . . . . . 5.1.1 Volume-Targeted Ventilation . . . . . . . 5.1.2 Pressure-Targeted Ventilation. . . . . . . 5.2 Setting the Respiratory Rate. . . . . . . . . . . . . . . 5.3 Setting the Flow Rate. . . . . . . . . . . . . . . . . . . . . 5.4 Setting the Ratio of Inspiration to Expiration (I:E Ratio) . . . . . . . . . . . . . . . . . . 5.5 Setting the Flow Profile . . . . . . . . . . . . . . . . . . . 5.5.1 The Square Waveform . . . . . . . . . . . . . 5.5.2 The Decelerating Waveform . . . . . . . . 5.5.3 The Accelerating Waveform . . . . . . . . 5.5.4 The Sine Waveform. . . . . . . . . . . . . . . . 5.6 Setting the Trigger Sensitivity . . . . . . . . . . . . . . 5.7 Setting PEEP. . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.7.1 Improvement in Oxygenation . . . . . . . 5.7.2 Protection Against Barotrauma and Lung Injury. . . . . . . . . . . . . . . . . . . 5.7.3 Overcoming Auto-PEEP . . . . . . . . . . . 5.8 Indications for PEEP . . . . . . . . . . . . . . . . . . . . . 5.9 Forms of PEEP . . . . . . . . . . . . . . . . . . . . . . . . . . 5.10 Titrating PEEP . . . . . . . . . . . . . . . . . . . . . . . . . . 5.10.1 Other Advantages of PEEP. . . . . . . . . 5.10.2 Disadvantages of PEEP . . . . . . . . . . . . 5.11 Optimizing Ventilator Settings for Better Oxygenation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.11.1 Increasing the FIO2 . . . . . . . . . . . . . . . .

xiii

102 102 103 103 106 107 108 109 110 115 115 115 116 117 118 119 122 122 123 123 123 123 124 124 125 126 127 127 127 130 131 131 131

xiv

Contents

5.11.2 Increasing the Alveolar Ventilation . . . . . . . . . . . . . . . . . . . . . . . 5.12 PEEP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.12.1 Flow Waveforms . . . . . . . . . . . . . . . . . . 5.12.2 Inspiratory Time . . . . . . . . . . . . . . . . . . 5.12.3 Inverse Ratio Ventilation. . . . . . . . . . . 5.12.4 Prone Ventilation . . . . . . . . . . . . . . . . . 5.12.5 Reducing Oxygen Consumption . . . . . 5.12.6 Increasing Oxygen Carrying Capacity . . . . . . . . . . . . . . . . . 5.12.7 Footnote . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

7

Ventilator Alarms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.1 Low Expired Minute Volume Alarm . . . . . . . . 6.2 High Expired Minute Volume Alarm. . . . . . . . 6.3 Upper Airway Pressure Limit Alarm . . . . . . . . 6.4 Low Airway Pressure Limit Alarm. . . . . . . . . . 6.5 Oxygen Concentration Alarms . . . . . . . . . . . . . 6.6 Low Oxygen Concentration (FIO2) Alarm . . . 6.7 Upper Oxygen Concentration (FIO2) Alarm . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.8 Power Failure. . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.9 Apnea Alarm. . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.10 Two-Minute Button . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Monitoring Gas Exchange in the Mechanically Ventilated Patient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.1 The Arterial Oxygen Tension . . . . . . . . . . . . . . 7.2 Pulse Oximetry . . . . . . . . . . . . . . . . . . . . . . . . . . 7.2.1 Principle of Pulse Oximetry. . . . . . . . . 7.3 Transcutaneous Blood Gas Monitoring . . . . . . 7.4 Monitoring Tissue Oxygenation . . . . . . . . . . . . 7.4.1 Oxygen Extraction Ratio and DO2 crit . . . . . . . . . . . . . . . . . . . . . . . 7.5 Capnography . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

132 132 132 133 133 134 134 134 135 136 141 141 143 144 146 146 146 147 147 147 148 148

149 149 156 160 169 171 172 175 183

Contents

8

9

Monitoring Lung Mechanics in the Mechanically Ventilated Patient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.1 Ventilator Waveforms. . . . . . . . . . . . . . . . . . . . . 8.2 Scalars. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.2.1 The Pressure–Time scalar . . . . . . . . . . 8.2.2 Flow-Time Scalar . . . . . . . . . . . . . . . . . 8.2.3 Volume–Time Scalar. . . . . . . . . . . . . . . 8.3 The Loops . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3.1 Pressure–Volume Loop . . . . . . . . . . . . 8.3.2 The Flow–Volume Loop . . . . . . . . . . . 8.4 Patient-Ventilator Asynchrony . . . . . . . . . . . . . 8.4.1 Level of Ventilator Support and Work of Breathing. . . . . . . . . . . . . 8.4.2 Complete Support. . . . . . . . . . . . . . . . . 8.4.3 Partial Support . . . . . . . . . . . . . . . . . . . 8.4.4 Patient-Ventilator Asynchrony . . . . . . 8.4.5 Triggering Asynchrony . . . . . . . . . . . . . 8.4.6 Flow Asynchrony . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Mechanical Ventilation in Specific Disorders. . . . . . . 9.1 Myocardial Ischemia . . . . . . . . . . . . . . . . . . . . . 9.2 Hypovolemic Shock . . . . . . . . . . . . . . . . . . . . . . 9.3 Neurological Injury. . . . . . . . . . . . . . . . . . . . . . . 9.4 Acute Respiratory Distress Syndrome (ARDS). . . . . . . . . . . . . . . . . . . . . . . 9.4.1 Primary and Secondary ARDS . . . . . . 9.4.2 Pathophysiology . . . . . . . . . . . . . . . . . . 9.4.3 Ventilatory Strategies . . . . . . . . . . . . . . 9.5 Obstructive Lung Disease . . . . . . . . . . . . . . . . . 9.5.1 PaCO2 . . . . . . . . . . . . . . . . . . . . . . . . . . . 9.5.2 Modes of Ventilation in Obstructed Patients . . . . . . . . . . . . . 9.5.3 Ventilator Settings in Airflow Obstruction . . . . . . . . . . . . . . . . . . . . . . 9.5.4 Bronchopleural Fistula. . . . . . . . . . . . . 9.6 Neuromuscular Disease . . . . . . . . . . . . . . . . . . . 9.6.1 Lung Function . . . . . . . . . . . . . . . . . . . .

xv

189 189 190 190 196 200 203 203 215 223 223 224 224 225 226 227 238 241 241 244 245 248 249 250 252 266 268 269 272 278 279 282

xvi

Contents

9.6.2

Inspiratory Muscle Recruitment in Neuromuscular Disease . . . . . . . . . . 9.6.3 Expiratory Muscle Recruitment in Neuromuscular Disease . . . . . . . . . . 9.6.4 Bulbar Muscles Involvement in Neuromuscular Disease . . . . . . . . . . 9.6.5 Assessment of Lung Function . . . . . . . 9.6.6 Mechanical Ventilation in Neuromuscular Disease . . . . . . . . . . 9.7 Nonhomogenous Lung Disease . . . . . . . . . . . . 9.8 Mechanical Ventilation in Flail Chest . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 The Complications of Mechanical Ventilation. . . . . . 10.1 Peri-Intubation Complications . . . . . . . . . . . . . 10.1.1 Laryngeal Trauma . . . . . . . . . . . . . . . . . 10.1.2 Pharyngeal Trauma . . . . . . . . . . . . . . . . 10.1.3 Tracheal or Bronchial Rupture . . . . . . 10.1.4 Epistaxis . . . . . . . . . . . . . . . . . . . . . . . . . 10.1.5 Tooth Trauma . . . . . . . . . . . . . . . . . . . . 10.1.6 Cervical Spine Injury . . . . . . . . . . . . . . 10.1.7 Esophageal Intubation . . . . . . . . . . . . . 10.1.8 Esophageal Perforation . . . . . . . . . . . . 10.1.9 Right Main Bronchial Intubation . . . . 10.1.10 Arrhythmias . . . . . . . . . . . . . . . . . . . . . 10.1.11 Aspiration . . . . . . . . . . . . . . . . . . . . . . . 10.1.12 Bronchospasm . . . . . . . . . . . . . . . . . . . . 10.1.13 Neurologic Complications . . . . . . . . . . 10.2 Problems Occurring Acutely at any Stage . . . . 10.2.1 Endotracheal Tube Obstruction . . . . . 10.2.2 Airway Drying. . . . . . . . . . . . . . . . . . . . 10.2.3 Upward Migration of the Endotracheal Tube. . . . . . . . . . . 10.2.4 Self-Extubation . . . . . . . . . . . . . . . . . . . 10.2.5 Cuff Leak . . . . . . . . . . . . . . . . . . . . . . . . 10.2.6 Ventilator-Associated Lung Injury (VALI) and Ventilator-Induced Lung Injury (VILI) . . . . . . . . . . . . . . . . 10.3 Delayed Complications (Fig. 10.5) . . . . . . . . . . 10.3.1 Sinusitis . . . . . . . . . . . . . . . . . . . . . . . . .

284 284 285 286 286 288 289 290 305 306 306 306 307 307 308 308 309 310 310 311 312 312 312 312 313 313 314 314 315

318 322 322

Contents

10.3.2 Tracheoesophageal Fistula. . . . . . . . . . 10.3.3 Tracheoinnominate Artery Fistula . . . . . . . . . . . . . . . . . . . . 10.3.4 Tracheocutaneous Fistula. . . . . . . . . . . 10.4 Oxygen-Related Lung Complications . . . . . . . 10.4.1 Tracheobronchitis . . . . . . . . . . . . . . . . . 10.4.2 Adsorptive Altelectasis . . . . . . . . . . . . 10.4.3 Hyperoxic Hypercarbia . . . . . . . . . . . . 10.4.4 Diffuse Alveolar Damage . . . . . . . . . . 10.4.5 Bronchopulmonary Dysplasia. . . . . . . 10.4.6 Ventilator-Associated Pneumonia . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 Ventilator-Associated Pneumonia. . . . . . . . . . . . . . . . 11.1 Incidence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.2 Microbiology . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.3 Risk Factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.3.1 The Physical Effect of the Endotracheal Tube. . . . . . . . . . . 11.3.2 Alteration of Mucus Properties . . . . . 11.3.3 Microaspiration . . . . . . . . . . . . . . . . . . . 11.3.4 Biofilms . . . . . . . . . . . . . . . . . . . . . . . . . 11.3.5 Ventilator Tubings. . . . . . . . . . . . . . . . . 11.3.6 Gastric Feeds . . . . . . . . . . . . . . . . . . . . . 11.3.7 Sinusitis . . . . . . . . . . . . . . . . . . . . . . . . . 11.3.8 Respiratory Therapy Equipment . . . . 11.4 Position. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.5 Diagnosis of VAP . . . . . . . . . . . . . . . . . . . . . . . . 11.5.1 Sampling Methods . . . . . . . . . . . . . . . . 11.5.2 Interpretation of the Sample . . . . . . . . 11.6 Prevention of NP/VAP. . . . . . . . . . . . . . . . . . . . 11.6.1 Hand-Washing. . . . . . . . . . . . . . . . . . . . 11.6.2 Feeding and Nutrition . . . . . . . . . . . . . 11.6.3 Stress Ulcer Prophylaxis . . . . . . . . . . . 11.6.4 Topical Antibiotics . . . . . . . . . . . . . . . . 11.7 Interventions Related to the Endotracheal Tube and Ventilator Circuit . . . . . . . . . . . . . . . . 11.8 Treatment of Nosocomial Sinusitis . . . . . . . . . . 11.9 Treatment. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.9.1 Antibiotic Resistance . . . . . . . . . . . . . .

xvii

323 325 326 327 328 328 329 332 333 333 334 343 345 345 347 347 348 349 349 350 351 352 354 354 355 357 358 360 360 361 362 362 363 364 365 365

xviii

Contents

11.9.2 Pharmacokinetics . . . . . . . . . . . . . . . . . 11.9.3 Duration of Therapy. . . . . . . . . . . . . . . 11.9.4 Lack of Response to Therapy . . . . . . . 11.9.5 Drug Cycling . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

368 371 373 374 376

12 Discontinuation of Mechanical Ventilation . . . . . . . . 12.1 Weaning Parameters. . . . . . . . . . . . . . . . . . . . . . 12.2 Parameters that Assess Adequacy of Oxygenation . . . . . . . . . . . . . . . . . . . . . . . . . . 12.2.1 The PaO2:FIO2 Ratio . . . . . . . . . . . . . . 12.2.2 The A-a DO2 Gradient. . . . . . . . . . . . . 12.2.3 The PaO2/PAO2 Ratio. . . . . . . . . . . . . . 12.3 Parameters that Assess Respiratory Muscle Performance . . . . . . . . . . . 12.3.1 PImax . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12.3.2 Vital Capacity . . . . . . . . . . . . . . . . . . . . 12.3.3 Minute Ventilation . . . . . . . . . . . . . . . . 12.3.4 Respiratory Rate. . . . . . . . . . . . . . . . . . 12.4 Parameters that Assess Central Respiratory Drive . . . . . . . . . . . . . . . . . 12.4.1 Airway Occlusion Pressure . . . . . . . . . 12.4.2 Mean Inspiratory Flow (Vt /Ti) . . . . . . 12.5 Respiratory System Compliance and Work of Breathing. . . . . . . . . . . . . . . . . . . . 12.5.1 Work of Breathing . . . . . . . . . . . . . . . . 12.5.2 Compliance of the Respiratory System . . . . . . . . . . . . . . . . . . . . . . . . . . 12.6 Integrative Indices . . . . . . . . . . . . . . . . . . . . . . . 12.6.1 Simplified Weaning Index (SWI) . . . . 12.7 Methods of Weaning. . . . . . . . . . . . . . . . . . . . . . 12.7.1 Trials of Spontaneous Breathing (T-Piece Weaning) . . . . . . . 12.7.2 Synchronized IMV . . . . . . . . . . . . . . . . 12.7.3 Pressure Support Ventilation (PSV) . . . . . . . . . . . . . . . . . 12.7.4 Noninvasive Positive Pressure Ventilation (NIPPV). . . . . . .

391 393 394 395 396 396 396 396 397 398 398 399 399 399 400 400 401 401 403 404 405 406 407 409

Contents

xix

12.7.5 Extubation . . . . . . . . . . . . . . . . . . . . . . . 409 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 411 13 Noninvasive Ventilation in Acute Respiratory Failure . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.1 NIV and CPAP . . . . . . . . . . . . . . . . . . . . . . . . . . 13.2 Mechanism of Action . . . . . . . . . . . . . . . . . . . . . 13.2.1 Interface . . . . . . . . . . . . . . . . . . . . . . . . . 13.2.2 Modes . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.2.3 Devices . . . . . . . . . . . . . . . . . . . . . . . . . . 13.2.4 Humidification with NIV (see also Chap. 15) . . . . . . . . . . . . . . . . 13.3 Air Leaks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.4 Indications for NIV. . . . . . . . . . . . . . . . . . . . . . . 13.4.1 Hypoxemic Respiratory Failure . . . . . 13.4.2 Hypercapnic Respitatory Failure . . . . 13.4.3 Miscellaneous Indications . . . . . . . . . . 13.4.4 Steps for the Initiation of NIV . . . . . . 13.4.5 Complications . . . . . . . . . . . . . . . . . . . . 13.4.6 Contraindications . . . . . . . . . . . . . . . . . 13.4.7 Outcomes . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 Negative Pressure Ventilation . . . . . . . . . . . . . . . . . . . 14.1 Tank Ventilator (Iron Lung) . . . . . . . . . . . . . . . 14.2 The Body Suit (Jacket Ventilator, Poncho-Wrap, Pulmo-Wrap) . . . . . . . . . . . . . . . 14.3 Chest: Shell (Cuirass) . . . . . . . . . . . . . . . . . . . . . 14.4 Modes of Negative Pressure Ventilation . . . . . 14.5 Drawbacks of NPV . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 Airway Humidification in the Mechanically Ventilated Patient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15.1 The Role of the Nasal Mucosa . . . . . . . . . . . . . 15.2 The Isothermic Saturation Boundary. . . . . . . . 15.3 The Effect of the Endotracheal Tube . . . . . . . . 15.3.1 Overheated Air . . . . . . . . . . . . . . . . . . .

415 415 415 418 420 421 422 422 424 424 426 427 428 429 432 432 433 441 442 442 443 444 445 446

449 449 449 450 451

xx

Contents

15.4 Heated Humidifiers . . . . . . . . . . . . . . . . . . . . . . 15.5 Heat-Moisture Exchangers (HMEs) . . . . . . . . 15.6 Airway Humidification During Noninvasive Ventilation . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 Aerosol Therapy in the Mechanically Ventilated Patient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.1 Terminology. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16.2 The Behavior of Particles. . . . . . . . . . . . . . . . . . 16.3 Devices for Aerosol Delivery . . . . . . . . . . . . . . 16.3.1 Jet Nebulizers (Syn: Pneumatic Nebulizers) . . . . . . . . 16.3.2 Ultrasonic Nebulizers. . . . . . . . . . . . . . 16.3.3 Vibrating Mesh Nebulizers (VMNs) . . . . . . . . . . . . . . . 16.3.4 Nebulization in the Ventilated Patient. . . . . . . . . . . . 16.3.5 Nebulization of Other Drugs. . . . . . . . 16.3.6 Pressurized Metered-Dose Inhalers (MDIs) . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 Nonconventional Modes and Adjunctive Therapies for Mechanical Ventilation. . . . . . . . . . . . . 17.1 High-Frequency Ventilation . . . . . . . . . . . . . . . 17.2 High-Frequency Positive Pressure Ventilation (HFPPV) . . . . . . . . . . . . . 17.3 High-Frequency Jet Ventilation (HFJV) . . . . . 17.4 High-Frequency Oscillatory Ventilation (HFOV) . . . . . . . . . . . . . . . . . . . . . . 17.5 High-Frequency Percussive Ventilation (HFPV) . . . . . . . . . . . . . . . . . . . . . . 17.6 Extracorporeal Life Support (ECLS) . . . . . . . 17.6.1 Extracorporeal Membrane Oxygenation (ECMO) . . . . . . . . . . . . . 17.6.2 Extracorporeal CO2 Removal . . . . . . . 17.6.3 Indications for ECLS . . . . . . . . . . . . . . 17.6.4 Contraindications to ECLS . . . . . . . . .

453 454 456 457

463 463 464 464 464 468 469 469 471 471 473

479 480 482 482 484 485 486 486 487 487 488

Contents

17.7 17.8 17.9 17.10

xxi

Nitric Oxide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Surfactant Therapy . . . . . . . . . . . . . . . . . . . . . . . Helium–Oxygen Mixtures . . . . . . . . . . . . . . . . . Liquid Ventilation . . . . . . . . . . . . . . . . . . . . . . . . 17.10.1 Total Liquid Ventilation . . . . . . . . . . . . 17.10.2 Partial Liquid Ventilation . . . . . . . . . . 17.11 NAVA. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17.12 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

488 491 493 494 496 496 497 497 498

18 Case Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.1 Case 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.2 Case 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.3 Case 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.4 Case 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.5 Case 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.6 Case 6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.7 Case 7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.8 Case 8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.9 Case 9 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.10 Case 10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.11 Case 11 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18.12 Case 12 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

505 505 508 510 511 512 513 516 517 518 520 522 523

Subject Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 527

Chapter 1 Historical Aspects of Mechanical Ventilation

As early as in the fifth century bc, Hippocrates, described a technique for the prevention of asphyxiation. In his work, “Treatise on Air,” Hippocrates stated, “One should introduce a cannula into the trachea along the jawbone so that air can be drawn into the lungs.” Hippocrates thus provided the first description of endotracheal intubation (ET).4,10 The first form of mechanical ventilator can probably be credited to Paracelsus, who in 1530 used fire-bellows fitted with a tube to pump air into the patient’s mouth. In 1653, Andreas Vesalius recognized that artificial respiration could be administered by tracheotomising a dog.24 In his classic, “De Humani Corporis Fabricia,” Vesalius stated, “But that life may … be restored to the animal, an opening must be attempted in the trunk of the trachea, in which a tube of reed or cane should be put; you will then blow into this so that the lung may rise again and the animal take in air… And also as I do this, and take care that the lung is inflated in intervals, the motion of the heart and arteries does not stop….” A hundred years later, Robert Hooke duplicated Vesalius’ experiments on a thoracotomised dog, and while insufflating air into an opening made into the animal’s trachea, observed that “the dog… capable of being kept alive by the reciprocal blowing up of his lungs with Bellows, and they suffered to subside, for the space of an hour or more, after his Thorax had been so displayed, and his Aspera arteria cut off just below the Epiglottis and bound upon the nose of the Bellows.”11 Hooke also made the important observation that it was not merely A. Hasan, Understanding Mechanical Ventilation, DOI: 10.1007/978-1-84882-869-8_1, © Springer-Verlag London Limited 2010

1

2

Chapter 1.  Historical Aspects of Mechanical Ventilation

the regular movement of the thorax that prevented asphyxia, but the maintenance of phasic airflow into the lungs. What was possibly the first successful instance of human resuscitation by mouth-to-mouth breathing was described in 1744 by John Fothergill in England. The use of bellows to resuscitate victims of near-drowning was described by the Royal Humane Society in the eighteenth century.20 The society, also known as the “Society for the Rescue of Drowned Persons” was constituted in 1767, but the development of fatal pneumothoraces produced by vigorous attempts at resuscitation led to subsequent abandonment of such techniques. John Hunter’s innovative double-bellows system (one bellow for blowing in fresh air, and another for drawing out the contaminated air) was adapted by the Society in 1782, and introduced a new concept into ventilatory care. In 1880, the endotracheal route was used, possibly for the first time, for cannulation of the trachea, and emerged as a realistic alternative to tracheotomy.14 Appreciation of the fact that life could be sustained by supporting the function of the lungs (and indeed the circulation) by external means led to the development of machines devised for this purpose. In 1838, Scottish physician John Dalziez described the first tank ventilator. In 1864 a body-tank ventilator was developed by Alfred Jones of Kentucky.9 The patient was seated inside an air-tight box which enclosed his body, neck downwards. Negative pressure generated within the apparatus produced inspiration, and expiration was aided by the cyclical generation of positive pressure at the end of each inspiratory breath. Jones took out a patent on his device which claimed that it could cure not only paralysis, neuralgia, asthma and bronchitis, but also rheumatism, dyspepsia, seminal weakness and deafness. Woillez’s hand-cranked “spirophore” (1876) and Egon Braun’s small wooden tank for the resuscitation of asphyxiated children followed. The former, the doctor operated by cranking a handle; the latter needed the treating physician to vigorously suck and blow into a tube attached to the box that enclosed the patient. In respect of Wilhelm Shwake’s pneumatic chamber, the patient himself could lend a hand by pulling and pushing against the bellows.

Historical Aspects of Mechanical Ventilation

3

In 1929, Philip Drinker, Louis Shaw, and Charles McKhann at the Department of Ventilation, Illumination, and Physiology, of the Harvard Medical School introduced what they termed “an apparatus for the prolonged administration of artificial respiration.”9 This team which included an engineer (Drinker), a physiologist (Shaw), and a physician (McKhann) saw the development of what was dubbed “the iron lung.” Drinker’s ventilator relied on the application of negative pressure to expand the chest, in a manner similar to Alfred Jones’ ventilator. The subject (at first a paralyzed cat, and then usually a patient of poliomyelitis) was laid within an air-tight iron tank. A padded collar around the patient’s neck provided a seal, and the pressure within the tank was rhythmically lowered by pumps or bellows. Access to the patient for nursing was understandably limited, though ports were provided for auscultation and monitoring.* Emerson, in 1931 in a variation upon this theme incorporated an apparatus with which it was possible to additionally deliver positive pressure breaths at the mouth; this made nursing easier. The patient could now be supported on positive pressure breaths alone, while the tank was opened periodically for nursing and examination. Toward the end of the nineteenth century, a ventilator functioning on a similar principle as the iron tank was independently developed by Ignaz von Hauke of Austria, Rudolf Eisenmenger of Vienna, and Alexander Graham Bell of the USA. Named so because of its similarity to the fifteenth century body armor, the “Cuirass” consisted of a breast plate and a back plate secured together to form an air-tight seal. Again, negative pressure generated by means of bellows (and during subsequent years, by a motor from a vacuum cleaner) provided the negative pressure to repetitively expand the thoracic cage and so move air in and out of the lungs. The Cuirass, by leaving the patient’s arms unencumbered, and by

A rich American financier’s son who developed poliomyelitis during a visit to China was transported back home in a Drinker-tank by a dozen caregivers which included seven Chinese nurses. He used the iron lung for more than two decades during which he married and fathered three children.

*

4

Chapter 1.  Historical Aspects of Mechanical Ventilation

causing less circulatory embarrassment, offered certain advantages over the tank respirator; in fact, Eisenmenger’s Cuirass was as much used for circulatory assistance during resuscitation as it was for artificial ventilation. Despite its advantages, the Cuirass proved to be somewhat less efficient than the tank respirator in providing mechanical assistance to breathing. During the earliest years of the twentieth century, advances in the field of thoracic surgery saw the design of a surgical chamber by Ferdinand Sauerbruch in 1904. This chamber functioned much on the same lines as the tank respirator except that the chamber included not only the patient’s torso, but the surgeon himself.4 Brauer reversed Sauerbruch’s principle of ventilation by enclosing only the patient’s head within a much smaller chamber which provided a positive pressure. In 1911, Drager designed his “Pulmotor,” a resuscitation unit which provided positive pressure inflation to the patient by means of a mask held upon the face. A tilted head position along with cricoid pressure (to prevent gastric insufflation of air) aided ventilation. The unit was powered by a compressed gas cylinder, and used by the fire and police departments for the resuscitation of victims.18 Negative pressure ventilators were extensively used during the polio epidemic that ravaged Los Angeles in 1948 and Scandinavia in 1952. During the Scandinavian epidemic, nearly three thousand polio-affected patients were treated in the Community Diseases Hospital of Copenhagen over a period of less than 6 months.16 The catastrophic mortality during the early days of the epidemic saw the use of the cuffed tracheostomy tube for the first time, in patients outside operating theaters. The polio epidemics in USA and Denmark saw the development and refinement of many of the principles of positive pressure ventilation. In 1950, responding to a need for better ventilators, Ray Bennet and colleagues developed an accessory attachment with which it became possible to intermittently administer positive pressure breaths in synchrony with the negative pressure breaths, delivered by a tank ventilator.3 The supplementation of negative pressure ventilation with intermittent

Historical Aspects of Mechanical Ventilation

5

positive pressure breaths did result in a substantial reduction in mortality.9,12,13 Bennet’s valve had originally been designed to enable pilots to breathe comfortably at high altitudes. The end of the Second World War saw the adaptation of the Bennet valve to regulate the flow of gases within mechanical ventilators.17 Likewise, Forrest Bird’s aviation experiences led to the design of the Bird Mark seven ventilator. Around this time, interest predictably focused on the physiological effects of mechanical ventilation. Courmand and then Maloney and Whittenberger made important observations on the hemodynamic effects of mechanical ventilation.15,17 By the mid 1950s, the concept of controlled mechanical ventilation had emerged. Engstrom’s paper, published in 1963, expostulated upon the clinical effects of prolonged controlled ventilation.7 In this landmark report, Engstrom stressed on the “complete substitution of the spontaneous ventilation of the patient by taking over both the ventilatory work and the control of the adequacy of ventilation” and so brought into definition, the concept of CMV. Engstrom developed ventilator models in which the minute volume requirements of the patient could be set. Setting the respiratory rate within a given minute ventilation determined the backup tidal volumes, and the overall effect was remarkably similar to the IMV mode in vogue today. Improvements in the design of the Bennet ventilators saw the emergence of the familiar Puritan-Bennet machines. The popularity of the Bennet and Bird ventilators in USA (both of which were pressure cycled) soon came to be rivaled by the development of volume-cycled piston-driven ventilators. These volume preset Emerson ventilators better guaranteed tidal volumes, and became recognized as potential anesthesia machines, as well as respiratory devices for long-term ventilatory support. Toward the end of the 1960s, with increasing challenges being presented during the treatment of critically ill patients on artificial ventilation, there arose a need for specialized areas for superior supportive care. During this period, a new disease entity came to be recognized, the Adult Respiratory Distress Syndrome, or the acute respiratory distress syndrome

6

Chapter 1.  Historical Aspects of Mechanical Ventilation

(ARDS) as it is known today. Physicians were confronted with rising demands for the supportive care of patients with this condition. The Respiratory Intensive Care Unit emerged as an important area for the treatment of critically ill patients requiring intensive monitoring. The use of positive end-­ expiratory pressure (PEEP) for the management of ARDS patients came into vogue, principally through Ashbaugh and Petty’s revival of Poulton and Barach’s concepts of the 1930s. A number of investigators staked claim to the development of the concept of PEEP, but controversy did not preclude its useful application.19,21 In 1971, Gregory et al applied continuous positive pressure to the care of neonates with the neonatal respiratory distress syndrome (NRDS) and showed that pediatric mechanical ventilation was possible. Several departures from the original theme of positive pressure ventilation followed, including the development of heroic measures for artificial support.1,5,8 Today’s ventilators have evolved from simple mechanical devices into highly complex microprocessor controlled systems which make for smoother patient-ventilator interaction. Such sophistication has, however, shifted the appreciation of the ventilator’s operational intricacies into the sphere of a new and now indispensable specialist – the biomedical engineer. Of late, resurgence in the popularity of noninvasive positive pressure breathing and the advent of high frequency positive pressure ventilation have further invigorated the area of mechanical ventilation; it also remains to be seen whether the promise of certain as yet unconventional modes of ventilation will be borne out in the near future.

References 1. Anderson HL, Steimle C, Shapiro M, et  al Extracorporeal life support for adult cardiorespoiratory failure. Surgery. 1993; 114:161 2. Ashbaugh DG, Bigelow DB, Petty TL, et  al Acute respiratory distress in adults. Lancet. 1967;2:319–323

References

7

3. Bennet VR, Bower AE, Dillon JB, Axelrod B. Investigation on care and treatment of poliomyelitis patients. Ann West Med Surg. 1950;4:561–582 4. Comroe JH. Retrospectorscope: Insights into Medical Discovery. Menlo park, CA: Von Gehr; 1977 5. Downs JB, Stock MC. Airway pressure release ventilation: a new concept in ventilatory support. Crit Care Med. 1987;15:459 6. Drinker P, Shaw LA. An apparatus for the prolonged administration of artificial respiration. 1. A design for adults and children. J Clin Invest. 1929;7:229–247 7. Engstrom CG. The clinical application of prolonged controlled ventilation. Acta Anasthesiol Scand [Suppl]. 1963;13: 1–52 8. Fort PF, Farmer C, Westerman J, et al High-frequency oscillatory ventilation for adult respiratory distress syndrome. Crit Care Med. 1997;25:937 9. Grenvik A, Eross B, Powner D. Historical survey of mechanical ventilation. Int Anesthesiol Clin. 1980;18:1–9 10. Heironimus TW. Mechanical Artificial Ventilation, Springfield, III, Charles C. Thomas; 1971 11. Hooke M. Of preserving animals alive by blowing through their lungs with bellows. Philo Trans R Soc. 1667;2:539–540 12. Ibsen B. The anesthetist’s view point on treatment of respiratory complications in polio during epidemic in Copenhagen. Proc R Soc Med. 1954;47:72–74 13. Laurie G. Ventilator users, home care and independent living: An historical perspective. In: Kutscher AH, Gilgoff I (eds). The Ventilator: Psychosocial and Medical aspects. New York Foundation of Thanatology, 2001; p147–151. 14. Macewen W. Clinical observations on the introduction of tracheal tubes by the mouth instead of performing tracheotomy or laryngotomy. Br Med J. 1880;2(122–124):163–165 15. Maloney JV, Whittenberger JL. Clinical implications of pressures used in the body respiration. Am J Med Sci. 1951;221: 425–430 16. Meyers RA. Mechanical support of respiration. Surg Clin North Am. 1974;54:1115 17. Motley HL, Cournand A, Werko L, et al Studies of intermittent positive pressure breathing as a means of administering artificial respiration in a man. JAMA. 1948;137:370–387 18. Mushin WI, et  al Automatic Ventilation of the Lungs. 2nd ed. Oxford, England: Blackwell Scientific; 1979

8

Chapter 1.  Historical Aspects of Mechanical Ventilation

19. Petty TL, Nett LM, Ashbaugh DG. Improvement in oxygenation in the adult respiratory distress syndrome by positive end expiratory pressure (PEEP). Respir Care. 1971;16:173–176 20. Randel-Baker L. History of thoracic anesthesia. In: Mushin WW, ed. Thoracic anesthesia. Philadelphia: FA Davis; 1963:598–661 21. Springer PR, Stevens PM. The influence of PEEP on survival of patients in respiratory failure. Am J Med. 1979;66:196–200 22. Standiford TJ, Morganroth ML. High-frequency ventilation. Chest. 1989;96:1380 23. Stock MC, Downs JB, Frolicher DA. Airway pressure release ventilation. Crit Care Med. 1987;15:462 24. Vesalius A. De humani corporis fabrica, Lib VII, cap. XIX De vivorum sectione nonulla, Basle, Operinus, 1543;658

Chapter 2 The Indications for Mechanical Ventilation

Apart from its supportive role in patients undergoing operative procedures, mechanical ventilatory support is indicated when spontaneous ventilation is inadequate for the sustenance of life. The word support bears emphasis, for mechanical ventilation is not a cure for the disease for which it is instituted: it is at best a form of support, offering time and rest to the patient until the underlying disease processes are resolved. Results with mechanical ventilation are consistently better when mechanical ventilatory support is initiated early and electively rather than in a crash situation. The indications for mechanical ventilation may be viewed as falling under several broad categories (Fig. 2.1).

2.1  Hypoxia Mechanical ventilation is often electively instituted when it is not possible to maintain an adequate oxygen saturation of hemoglobin. While optimization of tissue oxygenation is the goal, it is rarely possible to reliably assess the extent of tissue hypoxia. Instead, indices of blood oxygenation may rather need to be relied upon. Increasing the fraction of inspired oxygen (FIO2) indiscriminately in an attempt to improve oxygenation may unnecessarily subject the patient to the danger of oxygen toxicity (these concepts will be addressed at a later stage). Mechanical ventilation enables better ­control A. Hasan, Understanding Mechanical Ventilation, DOI: 10.1007/978-1-84882-869-8_2, © Springer-Verlag London Limited 2010

9

10

Chapter 2.  The Indications for Mechanical Ventilation Indications for intubation

Indications for ventilation

Need to secure airway Depressed sensorium

Hypoxia: acute hypoxemic respiratory failure

Depressed airway reflexes

Hypoventilation

Upper airway instability after trauma

Unacceptably high work of breathing

Decreased airway patency Need for sedation in the setting of poor airway control Imaging (CT, MRT) and transportation of an unstable patient

Hemodynamic compromise Cardiorespiratory arrest Refractory shock Raised intracranial pressure Flail chest

Figure 2.1.  Indications for intubation & ventilation.

of hypoxemia with relatively low inspired O2 concentrations, thereby diminishing the risk of oxygen toxicity.

2.2 Hypoventilation A major indication for mechanical ventilation is when the alveolar ventilation falls short of the patient’s requirements. Conditions that depress the respiratory center produce a decline in alveolar ventilation with a rise in arterial CO2 tension. A rising PaCO2 can also result from the hypoventilation that results when fatiguing respiratory muscles are unable to sustain ventilation, as in a patient who is expending considerable effort in moving air into stiffened or obstructed lungs. Under such circumstances, mechanical ventilation may be used to support gas exchange until the patient’s respiratory drive has been restored, or tired respiratory muscles rejuvenated, and the inciting pathology significantly resolved (Fig. 2.2).

2.3  Increased Work of Breathing

11

Hypoventilation results from decreased bulk flow in and out of the lungs Inspiration results in the bulk flow of air into the lungs, up to the level of the smallest bronchioles. Further progress of the gas molecules is by the mechanism of facilitated diffusion peripherally Disorders in which bulk flow to the lungs is compromised include

CNS depression e.g., Sedative agents Cerebrovascular accidents

Neuro-muscular disorders

Spinal cord or peripheral nerve disorders

Aminoglycosides

e.g.,

Paralysing agents

e.g.,

Spinal trauma

Steroid myopathy

Central sleep apnea

Amyotrophic lateral sclerosis

Myasthenia gravis

Metabolic alkalosis

Polio

Muscular dystrophies

Myxedema Hyperoxia (Hyperoxic hypoventilation)

Multiple sclerosis Guillian Barre syndrome Botulism

Dyselectrolyte mias

Disorders affecting the thoracic cage e.g., Kyphoscoliosis Flail chest Ankylosing spondylosis

Proximal airway (extrapulmonary airway) obstruction e.g., Tracheal obstruction by stenosis, tumor etc Epiglottitis Obstructive sleep apnea

Poor nutrition Respiratory muscle fatigue

Figure 2.2.  Causes of Hypoventilation.

2.3 Increased Work of Breathing Another major category where assisted ventilation is used is in those situations in which excessive work of breathing results in hemodynamic compromise. Here, even though gas exchange may not be actually impaired, the increased work of breathing because of either high airway resistance or poor lung compliance may impose a substantial burden on, for example, a compromised myocardium. When oxygen delivery to the tissues is compromised on account of impaired myocardial function, mechanical ventilation by resting the respiratory muscles can reduce the work of breathing. This reduces the oxygen consumption of the respiratory muscles and results in better perfusion of the myocardium itself.

12

Chapter 2.  The Indications for Mechanical Ventilation

2.4 Other Indications In addition to these major indications, mechanical ventilation may be of value in certain specific conditions. The vasoconstriction produced by deliberate hyperventilation can reduce the volume of the cerebral vascular compartment, helping to reduce raised intracranial pressures. In flail chest, mechanical ventilation can be used to provide internal stabilization of the thorax when multiple rib fractures compromise the integrity of the chest wall; in such cases, mechanical ventilation using positive end-expiratory pressure (PEEP) normalizes thoracic and lung mechanics, so that adequate gas exchange becomes possible. Where postoperative pain or neuromuscular disease limits lung expansion, mechanical ventilation can be employed to preserve a reasonable functional residual capacity within the lungs and prevent atelectasis. These issues have been specifically addressed in Chap. 9.

2.5 Criteria for Intubation and Ventilation While the prevailing criteria for defining the need for intubation and ventilation of a patient in respiratory failure have met general acceptance, these are largely intuitive and based upon the subjective assessment of a patient’s condition (Fig. 2.3 and Table 2.1). See also Chap. 12 . Objective criteria that are in current use are a forced expiratory volume in the first second (FEV1) of less than 10 mL/kg body weight and a forced vital capacity (FVC) of less than 15 mL/kg body weight, both of which indicate a poor ventilatory capability. Similarly, a respiratory rate higher than 35 breaths/min would mean an unacceptably high work of breathing and a substantial degree of respiratory distress, and is recognized as one of the criteria for intubation and ventilation. A PaCO2 in excess of 55 mmHg (especially if rising, and in the presence of acidemia) would likewise imply the onset of respiratory muscle fatigue. Except in habitual CO2 retainers, a PaCO2 of

2.5  Criteria for Intubation and Ventilation

13

B. In such a case a normal PaCO2 means that the CO2 has begun to rise back towards normal as a result of respiratory muscle fatigue

A. Hyperventilation results in PaCO2 wash out, producing respiratory alkalosis

Figure 2.3.  PaCO2 in status asthmaticus.

55 mmHg and over would normally reflect severe respiratory muscle dysfunction. Documented PaCO2 from an earlier stage of the patient’s present illness may have considerable bearing on the interpretation of subsequent PaCO2 levels (Fig. 2.3). For example, in an asthmatic patient in acute severe exacerbation, ­bronchospasm-induced hyperventilation can be expected to “wash out” the CO2 from the blood, producing respiratory alkalosis. If in such a patient, the blood gas analysis were to show a normal PaCO2 level, this would imply that the hypoventilation produced by respiratory muscle fatigue has allowed the PaCO2 to rise back to normal. It is important to realize here, that although the PaCO2 is now in the normal range, it is actually on its way up, and if this is not appreciated, neither the PaCO2 nor the patient will stay normal for very long. A supranormal PaCO2 in status asthmaticus should certainly be a cause of alarm and reinforce the need for mechanical ventilatory support. A PaO2 of less than 55–60 mmHg on 0.5 FIO2 or a widened A-a DO2 gradient (of 450 mmHg and beyond on 100% O2)

Normal range

5–8 mL/kg

12–20 breaths/ min 50–60 mL/kg

350–600 L/min

Tidal volume (V1)

Respiratory frequency (f) Forced expired volume at 1 s (FEV1) Peak expiratory flow

pH

7.35–7.45

35 breaths/ min 88%) FlO21.0 PEEP 34 If after reaching PEEP34cm H2O, no improvement in seen, revert PEEP to 24cm H2O

Figure 9.5.  Incremental increase in PEEP and FIO2 for optimization of oxygenation.7???

258

Chapter 9.  Mechanical Ventilation in Specific Disorders

Turn up the FIO2 to 1.0

Wait 15−20 min

Wait approximately 10 min

If the desired results have not been acheived, reapply CPAP at 35 cm H2O for 30−40 s

Apply CPAP at 30 cm H2O for 30−40 s

Wait 15−20 min If the desired results have not been acheived, reapply CPAP at 40 cm H2O for 30−40 s

Figure 9.6.  Recruitment by sustained inflation (CPAP) technique.

if the patient is hypovolemic.100,158 PEEP can, by elevating the right atrial pressure, produce intracardiac shunting across a patent foramen ovale.36 To ensure optimization of PEEP, and to protect against alveolar distension and depression of cardiac output, PEEP is added at a relatively low level (e.g., at 5 cm H2O) and then gradually boosted in small increments of 3–5  cm H2O at periodic intervals, until the desired level of oxygenation is achieved (Fig. 9.6).142 It is worth reemphasizing that the target for oxygenation of the blood is a PaO2 much above 60  mmHg. PaO2 much above this level will not serve to increase the oxygen delivery to the tissues any further, as at this level, the hemoglobin is almost completely saturated. The aim is to achieve at least this much PaO2 with as low an FIO2 as possible, certainly of 0.6 or less. It is hoped that this can be managed with a reasonable level of PEEP, but in ARDS, it is usual for a PEEP level of at least 15 cm H2O to be needed to bring this about.28 The construction of pressure–volume curves and the relevance of ventilating the patient at a level between the lower inflection point and the upper deflection point has been discussed in 5.10.

9.4  Acute Respiratory Distress Syndrome (ARDS)

259

9.4.3.8 Inspiratory Time Increasing the inspiratory time (Ti) can also help in decreasing the hypoxemia. Because of the heterogeneity of involvement in ARDS, alveoli with varying time constants are dispersed throughout the lungs: during inspiration, diseased alveoli take longer to fill up (see 3.5). With longer inspiratory time settings, diseased units are given the extra time they need to open up, and as a result, can now participate in ventilation. The inspiratory time can be increased by reducing the inspiratory flow rate. Inspiratory time can also be increased by using a decelerating waveform, wherein the inspiratory flow slows down progressively as inspiration advances. By adding an inspiratory pause, the lung can be held open at end-inspiration for the duration of the applied pause, achieving the same goal. Increasing the inspiratory time has been shown to improve blood oxygenation in ARDS.

9.4.3.9 Inverse Ratio Ventilation A mode of ventilation that has emerged as a logical extension of the above strategy is the inverse ratio ventilation (IRV) mode. Normally, inspiratory time occupies about a third of the respiratory cycle; two thirds of the respiratory cycle is spent in expiration. IRV is the reversal of this ratio; the inspiratory time is actually made to equal or exceed the expiratory time to enable improvement in oxygenation by the mechanisms just discussed. Controversies still surround the use of IRV.105,123 (see 5.4 & 5.12.3). It is also unclear whether the better oxygenation translates into improved survival.8 Owing to the prolongation of Ti (with the consequent decrease in expiration time) air trapping is possible, with its potential to cause barotrauma and hypotension in an already fragile situation. This problem usually occurs when the I:E ratio exceeds 2:1.106 Since the pattern of breathing in IRV is so different from the physiological pattern, patients generally tend to tolerate it poorly, and frequently require heavy sedation if not actual pharmacological paralysis.

260

Chapter 9.  Mechanical Ventilation in Specific Disorders

9.4.3.10 Recruitment Maneuvers When alveoli are atelectatic, the need for high pressures to achieve a given change in volume translates into poor lung compliance. Also, by repeatedly opening and closing during the respiratory cycle, unstable alveoli generate shear forces at their interface with relatively normal alveoli, increasing the probability of injury in these regions (atelectrauma).128,173 Lungs ventilated with low tidal volumes are likely to develop regional atelectasis. Since the dependent regions bear the weight of the lung above them, it is reasonable to suppose that some of the airspaces in these regions may be collapsed: therefore these should theoretically be recruitable (this concept is undoubtedly oversimplistic). Although PEEP is helpful in keeping the functioning lung units open, it cannot prize open up collapsed airspaces. Lung recruitment can be performed by several methods,80 but which of these is best, is currently uncertain. The most usually employed of these techniques is the sustained inflation (or CPAP) method.96,132 The incremental PEEP technique – used with pressure control ventilation – is generally used when there is a lack of response to sustained inflation. Lung recruitment by intermittent sighs (approximately 150% of tidal volume) may be effective in extrapulmonary ARDS or during prone ventilation (Fig. 9.7).136 Compared to the secondary (extrapulmonary) form, not only is the primary ARDS less responsive to recruitment, but is also more susceptible to hemodynamic compromise during the maneuver.97 Lung recruitment maneuvers are potentially harmful, and can cause severe hemodynamic compromise and barotrauma. They are contraindicated when the patient is hemodynamically unstable, or in the presence of intracranial hypertension, bronchospasm, lung bullae, or an untreated pneumothorax. Finally, not all patients respond to lung recruitment67; any improvements in hypoxemia may be transient. Still, despite their theoretical potential to overdistend the lung, recruitment maneuvers appear to be reasonably safe when used carefully. It should be kept in mind that the

9.4  Acute Respiratory Distress Syndrome (ARDS)

261

Patient Bronchial Small airways breathes at a lumen is tend to high lung a function of collapse volume in lung volume: during order to Expiratory Inspiratory expiratory expiration. increase lateral difficulty difficulty diameter is Expiratory air traction on less than trapping airway walls– inspiratory results and so prevent diameter airway collapse

Figure 9.7.  Pattern of breathing with dynamic hyperinflation.

pressures used in “cranking” the lung open are far in excess of those considered safe during tidal ventilation. The exercise should be aborted if the MAP falls to 140) or bradycardia (HR 100 mL returned is truly representative of one million-odd alveoli sampled during a typical broncho-alveolar lavage (BAL): the number of pathogens in 100 mL of the returning fluid will be about one million cfu. In contrast, colonizing organisms prevail at much lower concentrations (200 >0.35  frequency of respiration PIP > peak Inspiratory pressure PEEP > positive end-expiratory pressure MIP > maximal inspiratory pressure (the maximal negative pressure recorded during a 20-s occlusion of the airway.

 n SWI of >9 predicts weaning success 93% of the A time; an SWI of >11 is associated with weaning failure.

404

Chapter 12.  Discontinuation of Mechanical Ventilation

Box 12.3 The Compliance, Rate, Oxygenation, and Pressure (CROP) Index The CROP index factors in not only the demands on the respiratory system, but also the ability of the respiratory muscles to handle these demands. CROP index > [Cdyn × MIP × (PaO2/PAO2)]/f Cdyn > dynamic compliance MIP > maximal inspiratory pressure (the maximal negative pressure recorded during a 20-s occlusion of the airway. PaO2 > Oxygen tension of the arterial blood PAO2 > Oxygen tension of the alveolar air f > frequency of respiration.

 CROP index of >13 mL/breath per min generally A correlates with successful weaning: in one prospective study, its positive and negative predictive values were determined to be 71 and 70%, respectively47.

Box 12.4 Pressure-Time Product (PTI) The PTI is another integrative index: it is the time integral of respiratory muscle pressure. Since it quantifies the minute ventilation required to maintain the PaCO2 at 40  mmHg (VE40), it is essentially a measure of respiratory muscle endurance, in addition to being a measure of gas exchange.13

12.7 Methods of Weaning Weaning can either be accomplished by trials of spontaneous breathing on the endotracheal tube for progressively longer periods of time, or by gradually decreasing the level of support

12.7  Methods of Weaning

405

on IMV, SIMV+PS, or pressure support ventilation (PSV). As a lead-up to either method, the tidal volumes delivered by the ventilator should be reduced to the tidal volumes the patient is expected to generate when the mechanical ventilatory support is discontinued. Many COPD patients are expected to retain some amount of CO2 even when they are clinically stable, and their PaCO2 levels should be brought up slowly to their expected “baseline” before commencement of the weaning trial.

12.7.1 Trials of Spontaneous Breathing (T-Piece Weaning) During a T-piece trial, the patient is disconnected from the ventilator, a T-piece is attached to the endotracheal tube, and an appropriate concentration of O2 is administered through one limb of the T-piece. The patient is encouraged to breathe on his own through the endotracheal tube, initially for brief intervals of time. These periods of spontaneous breathing are progressively lengthened until the patient is capable of breathing on his own for a reasonable period of time without manifesting any signs of distress. What constitutes a “reasonable period of time” has not been determined, but clinical experience predicts weanability when the patient is able to comfortably tolerate spontaneous T-piece breathing for 1–24 h. Shorter trials of between 30 and 120 min of spontaneous breathing may be just as effective in predicting weaning success.7 Neither has it been resolved what intervals of “rest” on the ventilator are optimal between attempts at spontaneous breathing,6 but again clinical experience points to a range of 1–3 h as sufficient. It is possible that even as few as one trial of spontaneous breathing a day is sufficient. Trials of spontaneous breathing should be terminated immediately at any stage should any signs of cardiorespiratory distress develop (Fig. 12.2). Although relatively more labor intensive for the nursing staff and respiratory therapists, the T-piece method serves quite well in patients without significant lung disease. The danger lies in that prolonged trials of spontaneous breathing can deplete

406

Chapter 12.  Discontinuation of Mechanical Ventilation An increase in respiratoy rate to>40 breaths/min (with an increase in respiratory rate over the baseline respiratory rate by 10 breaths/min) Recruitment of the accessory muscles of respiration A rise or fall in heart rate by>20 beats/min New arrythmia Fall in PaO2 to less than 60 mmHg (or fall in SpO2to5 mmHg (or fall in pH to 8 on the glasgow coma Blotter method

Spirometric method

• ETT is disconnected from the ventilator circuit • An index card or a blotting paper is held about 1–2 cm away from the end of the ETT • The patient is instructed to cough forcefully • An ability to moisten the card with three out of four cough attempts correlates 3 times more strongly with successful extubation than does an inability to do so16

• A spirometer is introduced into the ventilator circuit • The patient is instructed to cough forcefully • The cough peak flow is measured on the spirometer • A cough peak flow of 60L/min or more correlates 5 times more strongly with successful extubation than does a cough peak flow  radius LaPlace’s Law states that the change in pressure within an air bubble is inversely proportional to its radius. Applied to the air unit, during exhalation, as the denominator in the equation (the radius of the alveolus) decreases, the change in pressure within will tend to collapse it – unless the numerator (the surface tension) shows a parallel decrease. In other words, an air unit will remain stable only when a reduction in its radius is paralleled by a reduction in the surface tension. Surface tension forces within the alveolus are dynamic and match the alveolar diameter: they can range from 40 dynes/ cm at end inspiration to 10 dynes/cm at end expiration.

17.9  Helium–Oxygen Mixtures

493

the inherent drawback that surfactant does not precisely aerosolize to its targeted site – the surfactant-depleted atelectatic alveoli. It may, therefore, be more effectual to directly instill surfactant boluses into the endotracheal tube62 or lavage the bronchopulmonary segments with surfactant through the bronchoscope.65 Postinstillation, adequate PEEP and tidal volumes must be applied to ensure the success of the treatment.38 Unfortunately, surfactant replacement therapy has not yet fulfilled the promise it initially showed, though with the advent of superior formulations not only ARDS but other diseases such as alveolar proteinosis,21 interstitial lung disease,42 asthma, and cystic fibrosis19 could be targeted as well.

17.9 Helium–Oxygen Mixtures As described in Chap. 3, airflow is turbulent in the larger airways and more laminar in the smaller. The flow of gases through straight and rigid tubes is governed by physical laws. At low velocities, particles travel in a streamlined manner, parallel to the sides of the tube: this is termed laminar flow. The gas front is parabolic, since the particles in the center of the tube move faster than those at the periphery. At high velocities, particles move in an unpredictable chaotic manner; this is termed turbulent flow. The critical mathematical parameter that determines that the flow is laminar or turbulent is a dimensionless number – the Reynold number. Reynold number is itself a function of the diameter of the tube as well as the density, velocity, and the viscosity of the gas. The last of these is relevant to the discussion that follows. Looking at the equation (Box 17.5) it is evident that in a given tube – or airway – when flow is constant, it is the density of the gas which determines whether the airflow is laminar or turbulent. In a straight, rigid, smooth nonbranching tube, a Reynold number below 2,000 results in laminar flow; a number above 4,000 determines turbulent flow. Unfortunately the airway is none of these, and as a result of considerable turbulence, the above equation, applied to the human airway, lacks

494

Chapter 17.  Nonconventional Modes

precision. It, nevertheless, follows that if gas with a lower density than air is breathed, turbulence can be reduced. Helium is such a gas, with a density of 0.8 g/L (that of air is 1.29  g/L). The addition of helium (in place of nitrogen) to oxygen lowers the Reynolds number of the gas mixture, and so reduces turbulence. Mixtures of helium and oxygen (heliox) in the proportion 80:20 or 70:30 are used in clinical practice to try to overcome the high airway pressure in refractory asthma and respiratory distress syndrome in children.8,37 The flow of gas through an orifice being inversely proportional to the square root of its density (Graham’s Law), Heliox will flow almost twice as fast than an oxygen flow-meter indicates.20 Unless the lower density of this gas mixture is factored in, errors can occur during the calculation of flows while ventilating patients with heliox. Similarly, since the set FIO2 may not reflect the fraction of oxygen actually delivered to the patient who is breathing heliox, oxygen analyzers must be relied upon while programming oxygen delivery on the ventilator.

Box 17.5 Determinants of the Reynold Number R > pdv/u where, R > Reynolds number p > density d > airway diameter v > velocity of gas u > viscosity of the gas

17.10 Liquid Ventilation Liquid ventilation involves the insufflation of lungs by perfluorochemicals rather than air for the purpose of gas exchange. Attempts to improve oxygen diffusion through

17.10  Liquid Ventilation

495

diseased lungs using hyperbaric saline ventilation41 led to the development of perfluorocarbons as media for facilitating gas exchange.12 When the hydrogen atoms of hydrocarbons are replaced by the halogen fluorine, biologically inert perfluorocarbons are formed. Perfluorocarbons are colorless but radio-opaque fluids. O2 and CO2 have a high solubility into these chemicals, which as a result, provide a large reservoir for these gases. The solubility of oxygen within perfluorocarbons exceeds by approximately 15-fold that within plasma.59 Perfluorocarbons have certain properties which make them well suited for their purpose as media for liquid ventilation. Unlike saline, they do not remove surfactant from alveoli.17 By abolishing the air–liquid interface at the surface of the alveolar epithelium, they substantially diminish surface tension especially in surfactant-depleted air units. By filling up alveoli, they prevent alveolar closure, recruit alveoli, and improve functional residual capacity. It is possible that the reason for improvement in oxygenation is alveolar recruitment. The gravitational descent of the dense liquid into the nether regions of the lung may also be responsible for the redistribution of perfusion to the relatively well-ventilated nondependent areas. By assuring alveolar patency even at relatively low airway pressures, perfluorocarbons can lower the risk for barotrauma. Inappropriately large boluses of perfluorocarbons can in fact distend the lung in a similar manner to PEEP, and are similarly capable of compromising the cardiac output. Since they are denser than water, perfluorocarbons gravitate to the dependent air units where the lung pathology is often most severe; mucus and inflammatory debris being lighter than perfluorocarbons float upward and can so be removed. The rinsing of the lung of inflammatory mediators may possibly modify neutrophil function and help limit lung inflammation and injury.61 Owing to their higher heat capacity, perfluorocarbons can be utilized to regulate core body temperature.25,58 Perfluorocarbons are immiscible with body fluids and negligible quantities diffuse into pulmonary capillary blood.67

496

Chapter 17.  Nonconventional Modes

17.10.1 Total Liquid Ventilation During total liquid lung ventilation (TLV), the lungs are completely filled up with oxygenated perfluorocarbons and tidal boluses of perfluorocarbons are periodically pumped in and out of the lungs. This involves the use of complex equipment for the transportation of these dense fluids to and from the lungs and for the extracorporeal oxygenation and CO2 removal from the fluid. The repeated instillation into and removal of perfluorocarbons from the lungs also serves to rinse the alveoli.54

17.10.2 Partial Liquid Ventilation Technically less demanding, partial liquid ventilation involves the instillation of perfluorocarbons that are quantitatively equivalent to the functional residual capacity of the lungs. In this way, the partially filled lungs can be ventilated by a conventional mechanical ventilator rather than the complex equipment used for TLV. Experimentally, partial liquid lung ventilation has been convincingly shown to improve lung compliance, reduce the shunt fraction, and improve gas exchange.1,47 As in the case of TLV, inflammatory debris from the distal parts of the lung is mobilized proximally; unlike in TLV, since fluid exchanges are not carried out, exudates have to be suctioned out of the proximal airways.34,36 Indications: Potential applications for liquid ventilation include neonatal hyaline membrane disease,54 meconium aspiration,35 and persistent primary pulmonary hypertension of the newborn.13 Although the role of liquid ventilation in adults appears to be most relevant in ARDS, its use has also been proposed in pneumonia where the lavaging of infected lungs with perfluorocarbons enables purging of bacteria and exudates,35 and even facilitates the delivery of antibiotics suspended within the perfluorocarbon directly to the infected alveoli.15 Perfluoro­ carbons have also been utilized for the prevention of lung

17.12  Conclusion

497

injury during cardiopulmonary bypass,10 and for conservation donor lungs before transplantation.68 The potential side effects – including bleeding, mucous plug formation, and pneumothorax – are inseparable from those of the inciting pathology.34 Because they are radio-opaque, perfluorocarbons dramatically opacify the lungs; needless to say, their use negates the diagnostic value of the chest radiograph. Uncontrolled trials in humans have certainly shown reason for optimism, though the specific role of liquid ventilation yet remains to be defined.

17.11 NAVA Neurally adjusted ventilatory assist (NAVA)60 is a novel closed-loop mode of ventilation that was specifically designed to counteract patient-ventilator asynchrony. The delay between the neural trigger and the ventilator response time is an important cause of patient-ventilator asynchrony. At the present time, the diaphragm is the most proximal level at which a neural trigger signal can be electrically sensed. In NAVA, an array of nasogastric tube-mounted electrodes placed at the level of the diaphragm detects the intensity of the patient’s neural output, and the strength of this signal, integrated to the ventilator output, determines support provided by the ventilator. The ventilator cycles on at the onset of neural inspiration and cycles off when neural expiration begins. Since the patient interacts intimately with the machine, in theory at least, synchronization is better.

17.12 Conclusion As in any other field in clinical medicine, mechanical ventilation is not an exact science. Much conceptual change has occurred during the past three decades and many of the strategies of ventilation will continue to change. It is almost certain that the near future will see a major revolution in this field, which should fill up the deficiencies that exist in current practice.

498

Chapter 17.  Nonconventional Modes

References 1. Al-Rahmani A, Awad K, Miller TF, Wolfson MR, Shaffer TH. Effects of partial liquid ventilation with perfluorodecalin in the juvenile rabbit lung after saline injury. Crit Care Med. 2000;28:1459–1464 2. Anderson HL, Steimle C, Shapiro M, et  al Extracorporeal life support for adult cardiorespiratory failure. Surgery. 1993;114:161 3. Bartlett RH. Extracorporeal life support for cardiopulmonary failure. Curr Probl Surg. 1990;27:621 4. Bernhard W, Haslam PL, Floros J. From birds to humans: new concepts on airways relative to alveolar surfactant. Am J Respir Cell Mol Biol. 2004;30:6–11 5. Bishop MJ, Benson MS, Sato P, et  al Comparison of high frequency ventilation with conventional mechanical ventilation for bronchopleural fistula. Anesth Analg. 1987;66:833 6. Bollen CW, van Well GT, Sherry T, et al High frequency oscillatory ventilation compared with conventional mechanical ventilation in adult respiratory distress syndrome: a randomized controlled trial. Crit Care. 2005;9:R430 7. Bone RC, Slotman G, Maunder R, et al Randomised double-blind, multicenter study of prostaglandin E1 in patients with the adult respiratory distress syndrome. Prostaglandin E1 Study Group. Chest. 1989;96:114 8. Boorstein JM, Boorstein SM, Humphries GN, Johnston CC. Using helium-oxygen mixtures in the emergency management of acute upper airway obstruction. Ann Emerg Med. 1989;18:688 9. Brunet F, Belghith M, Mira JP, et al Extracorporeal carbon dioxide removal and low-frequency positive-pressure ventilation. Improvement in arterial oxygenation with reduction of risk of pulmonary barotrauma in patients with adult respiratory distress syndrome. Chest. 1993;104:889 10. Cannon ML, Cheifetz IM, Craig DM, et  al Optimizing liquid ventilation as a lung protection strategy for neonatal cardiopulmonary bypass: full functional residual capacity dosing is more effective than half functional residual capacity dosing. Crit Care Med. 1999;27:1140 11. Whittaker KB. Neonatal mechanical ventilation In: Chang DW. Clinical Applications of Mechanical Ventilation. 3rd  ed. Delmar learning; 2006;541–549 12. Clark JC Jr, Gollan F. Survival of mammals breathing organic liquids equilibrated with oxygen at atmospheric pressure. Science. 1966;152:1755

References

499

13. Cox CA, Wolfson MR, Shaffer TH. Liquid ventilation: a comprehensive overview. Neonatal Netw. 1996;15(3):31–43 14. Cueto E, Lopez-Herce J, Sanchez A, et  al Life-threatening effects of discountinuing inhaled nitric oxide in children. Acta Paediatr. 1997;86:1337–1339 15. Cullen AB, Cox CA, Hip SJ, et al Intra-tracheal delivery strategy of gentamicin with partial liquid ventilation. Respir Med. 1999; 93:770 16. David M, Weiler N, Heinrichs W, et al High-frequency oscillatory ventilation in adult acute respiratory distress syndrome. Intensive Care Med. 2003;29:1656 17. Degraeuwe PLJ, Vos GD, Blanco CE. Perfluorochemical liquid ventilation: from the animal laboratory to the intensive care unit. Int J Artif Organs. 1995;18:674 18. Demory D, Michelet P, Arnal JM, et al High-frequency oscillatory ventilation following prone positioning prevents a further impairment in oxygenation. Crit Care Med. 2007;35:106 19. DeSanctis GT, Tomkiewicz RP, Rubin BK, et al Exogenous surfactant enhances mucociliary clearance in the anesthetized dog. Eur Respir J. 1994;7:1616–1621 20. Devabhaktuni VG, Torres A Jr, Wilson S, Yeh MP. Effect of nitric oxide, perfluorocarbon, and heliox on minute volume measurement and ventilator volumes delivered. Crit Care Med. 1999;27:1603 21. Doyle IR, Davidson KG, Barr HA, et al Quantity and structure of surfactant proteins vary among patients with alveolar proteinosis. Am J Respir Crit Care Med. 1998;157:658–664 22. Eastman A, Holland D, Higgins J, et al High-frequency percussive ventilation improves oxygenation in trauma patients with acute respiratory distress syndrome: a retrospective review. Am J Surg. 2006;192:191 23. Ferguson ND, Chiche JD, Kacmarek RM, et al Combining highfrequency oscillatory ventilation and recruitment maneuvers in adults with early acute respiratory distress syndrome: the treatment with oscillation and an open lung strategy (TOOLS) Trial pilot study. Crit Care Med. 2005;33:479 24. Fessler HE, Derdak S, Ferguson ND, et al A protocol for highfrequency oscillatory ventilation in adults: results from a roundtable discussion. Crit Care Med. 2007;35:1649 25. Forman DL, et  al A new approach to induced hypothermia. J Surg Res. 1986;40:36 26. Froese AB. Unconventional methods of ventilator support: high frequency ventilation. In: Tobin MJ, ed. Principles and Practice of Mechanical Ventilation. McGraw-Hill; 2006

500

Chapter 17.  Nonconventional Modes

27. Frumin MJ, Epstein RM, Cohen G. Apneic oxygenation in man. Anesthesiology. 1959;20:789–798 28. Gattinoni L, Kolobow T, Tomlinson T, et al Low-frequency positive pressure ventilation with extracorporeal carbon dioxide removal (LFPPV-ECCO2R): an experimental study. Anesth Analg. 1978;57:470–477 29. Gerlach H, Rossaint R, Pappert D, et al Time-course and doseresponse of nitric oxide inhalation for systemic oxygenation and pulmonary hypertension in patients with adult respiratory distress syndrome. Eur J Clin Invest. 1993;23:499–502 30. Gluck E, Heard S, Patel C, et al Use of ultrahigh frequency ventilation in patients with ARDS: a preliminary report. Chest. 1993;103:1413 31. Greenbaum R, Bay J, Hargreaves MD, et  al Effects of higher oxides of nitrogen on the anaesthetized dog. Br J Anaesth. 1967;39: 393–404 32. Hager DN, Fessler HE, Kaczka DW, et al Tidal volume delivery during high-frequency oscillatory ventilation in adults with acute respiratory distress syndrome. Crit Care Med. 2007;35:1522 33. Hemmila MR, Rowe SA, Boules TN, et  al Extracorporeal life support for sever acute respiratory distress syndrome in adults. Ann Surg. 2004;240:595 34. Hirschl RB. Advances in the management of respiratory failure. Liquid ventilation in the setting of respiratory failure. ASAIO J. 1996;42:20 35. Hirschl RB, Pranikoff T, Gauger P, et  al Liquid ventilation in adults, children and full-term neonates. Lancet. 1995;346:1201 36. Hirschl RB, Pranikoff T, Wise C, et  al Initial experience with partial liquid ventilation in adult patients with the acute respiratory distress syndrome. JAMA. 1996;275:383 37. Hohlfeld J, Fabel H, Hamm H. The role of pulmonary surfactant in obstructive airway disease. Eur Respir J. 1997;10:482–491 38. Ito Y, Manwell SEE, Kerr CL, et al Effect of ventilation strategies on the efficacy of exogenous surfactant therapy in a rabbit model of acute lung injury. Am J Respir Crit Care Med. 1998;157: 149–155 39. Kobayashi T, Nitta K, Ganzuka M, et al Inactivation of exogenous surfactant by pulmonary edema fluid. Pediatr Res. 1991;29:353–356 40. Kolla S, Lee WA, Hirschl RB, Bartlett RH. Extracorporeal life support for cardiovascular support in adults. ASAIO J. 1996;42:M809 41. Kylstra JA, Tissing MO, Van der Maen A. Of mice as fish. ASAIO Trans. 1962;8:378

References

501

42. McCormack FX, King TEJ, Voelker DR, et al Idiopathic pulmonary fibrosis. Abnormalities in the bronchoalveolar lavage content of surfactant protein A. Am Rev Respir Dis. 1991;144:160–166 43. Mehta S, MacDonald R, Hallett DC, et  al Acute oxygenation response to inhaled nitric oxide when combined with high-frequency oscillatory ventilation in adults with acute respiratory distress syndrome. Crit Care Med. 2003;31:383 44. Mehta S, Granton J, MacDonald RJ, et al High-frequency oscillatory ventilation in adults: the Toronto experience. Chest. 2004;126:518 45. Mentzelopoulos SD, Roussos C, Koutsoukou A, et  al Acute effects of combined high-frequency oscillation and tracheal gas insufflation in severe acute respiratory distress syndrome. Crit Care Med. 2007;35:1500 46. Milner AD, Hoskins EW. High frequency positive pressure ventilation in neonates. Arch Dis Child. 1989;64(1):1–3; [Fetal Neonatal ed] 47. Papo MC,Paczan PR,Fuhrman BP,et al Perfluorocarbonassociated gas exchange improves oxygenation, lung mechanics, and survival in a model of adult respiratory distress syndrome. Crit Care Med. 1996;24:466–474 48. Peek GJ, Clemens F, Elbourne D, et  al CESAR: conventional ventilatory support vs extracorporeal membrane oxygenation for severe adult respiratory failure. BMC Health Serv Res. 2006; 6:163 49. Putensen C, Rasanen J, Lopez FA, et al Continuous positive air way pressure modulates effect of inhaled nitric oxide on the ventilation- perfusion distributions in canine lung injury. Chest. 1994;106:1563–1569 50. Puybasset L, Rouby JJ, Mourgeon E, et  al Factors influencing cardiopulmonary effects of inhaled nitric oxide in acute respiratory failure. Am J Respir Crit Care Med. 1995;152:318–328 51. Quezado ZM, Natanson C, Karzai W, et  al Cardiopulmonary effects of inhaled nitric oxide in normal dogs and during E. coli pneumonia and sepsis. J Appl Physiol. 1998;84:107–115 52. Reper P, Wibaux O, Van Laeke P, et al High frequency percussive ventilation and conventional ventilation after smoke inhalation: a randomised study. Burns. 2002;28:503 53. Rialp G, Betbese AJ, Peres-Marquez M, et al Short-term effects of inhaled nitric oxide and prone position in pulmonary and extrapulmonary acute respiratory distress syndrome. Am J Respir Crit Care Med. 2001;164:243–249

502

Chapter 17.  Nonconventional Modes

54. Richman PS, Wolfson MR, Shaffer TH. Lung lavage with oxygenated perfluorochemical liquid in acute lung injury. Crit Care Med. 1993;21:768 55. Rossaint R, Falke KJ, Lopez F, et al Inhaled nitric oxide for the adult respiratory distress syndrome. N Engl J Med. 1993;328: 399–405 56. Salim A, Martin M. High-frequency percussive ventilation. Crit Care Med. 2005;33:S241 57. Seeger W, Grube C, Gunther A, Schmidt R. Surfactant inhibition by plasma proteins: differential sensitivity of various surfactant preparations. Eur Respir J. 1993;6:971–977 58. Shaffer TH, Forman D, Wolfson MR. The physiological effects of breathing fluorocarbon liquids at various temperatures. Undersea Biomed Res. 1984;11:287 59. Shaffer TH, Wolfson MR, Clark LC. State of the art review. Liquid ventilation. Pediatr Pulmonol. 1992;14:102 60. Sinderby C, Navalesi P, Beck J, et al Neural control of mechanical ventilation in respiratory failure. Nat Med. 1999;5:1433–1436 61. Smith TM, Steinhorn DM, Thusu K, et  al A liquid perfluorochemical decreases the in  vitro production of reactive oxygen species by alveolar macrophages. Crit Care Med. 1995;23:1533 62. Spragg R, Lewis J, Walmrath H, et al Effect of recombinant surfactant protein C based surfactant on patients with the acute respiratory distress syndrome. N Engl J Med. 2004;351:884–892 63. Veldhuizen R, Nag K, Orgeig S, et al The role of lipids in pulmonary surfactant. Biochim Biophys Acta. 1998;1408:90–108 64. Veldhuizen RA, Yao L, Lewis JF. An examination of the different variables affecting surfactant aggregate conversion in vitro. Exp Lung Res. 1999;25:127–141 65. Walmarth D, Grimminger F, Pappert D, et  al Bronchoscopic administration of bovine natural surfactant in ARDS and septic shock; impact on gas exchange and haemodynamics. Eur Respir J. 2002;19:805–810 66. Whitsett JA. Surfactant proteins in innate host defense of the lung. Biol Neonate. 2005;88:175–180 67. Wolfson MR, Kechner NE, Rubenstein D, et al Perfluorochemical (PFC) uptake and biodistribution following liquid assisted ventilation in the immature lamb. Pediatr Res. 1994;35:A246 68. Yoshida S, Sekine Y, Shinozuka N, et  al The efficacy of partial liquid ventilation in lung protection during hypotension and cardiac arrest: preliminary study of lung transplantation using non-heart-beating donors. J Heart Lung Transplant. 2005;24:723

References

503

69. Zapol WM, Snider MT, Hill JD, et al Extracorporeal membrane oxygenation in severe acute respiratory failure. A randomized prospective study. JAMA. 1979;242:2193 70. Zobel G, Dacar D, Rodl S. Proximal and tracheal airway pressure during different modes of mechanical ventilation: an animal model study. Pediatr Pulmonol. 1994;18:239

Chapter 18 Case Studies

The following modules represent some of the common situations that require troubleshooting on mechanically ventilated patients.

18.1 Case 1 Mr. A was ventilated 4 days ago for type-2 respiratory failure during a severe exacerbation of COPD. When recovering on the ventilator, he suddenly developed respiratory distress and tachycardia. SpO2 which had hitherto been stable at 92% on an FIO2 of 0.28, dropped to 69%. The peak airway pressure alarm became activated simultaneously with the onset of Mr. A’s respiratory distress, and tidal volumes (set at 480 mL on the assistcontrol mode) became severely pressure limited and dropped substantially. Appropriate action at this stage would be: (a) Disconnection from ventilator followed by bagging and suctioning the ET tube (b) Chest X-ray (c) ABG (d) A careful physical examination The correct answer is (a) since it permits immediate identification of whether the problem is with the machine or with the patient; it further enables diagnosis and treatment of an endotracheal tube block which is not only common, but must be urgently identified. A. Hasan, Understanding Mechanical Ventilation, DOI 10.1007/978-1-84882-869-8_18, © Springer-Verlag London Limited 2010

505

506

Chapter 18.  Case Studies

Mr. A was disconnected from the ventilator and manually ventilated with a bag and mask. The bag appeared stiff and substantial resistance to manual compression of the bag was encountered. A suction catheter passed down the endotracheal tube encountered significant resistance to its passage. After thorough suctioning, which resulted in the removal of sticky mucus, the suction catheter could now pass unimpeded down the endotracheal tube. Mr. A was reconnected to the ventilator and was now more comfortable; although the peak pressure was now lower, it was still high than before. Reasons for the persistently elevated peak pressure could include: (a) A partial blockage of endotracheal tube (b) Bronchospasm (c) A small “occult” pneumothorax (d) Any of the above The correct answer is (d), since all three conditions can elevate the peak airway pressure. Clinical examination did not reveal an overt bronchospasm, and a bedside chest film did not reveal a pneumothorax. The next logical option would be: (a) Arterial blood gas analysis (b) Change the endotracheal tube regardless of the fact that the suction catheter can be negotiated through it (c) Chest physiotherapy (d) Work up for pulmonary embolism The correct answer is (b). Even slight narrowing of the lumen of the endotracheal tube can result in significant airflow obstruction. The endotracheal tube was, a short time ago, completely blocked, and it is likely that encrustations remain, causing partial blockage. Apart from the risk of reblockage, breathing through a narrowed endotracheal tube entails a high work of breathing and predisposes to respiratory muscle fatigue. When in doubt, the endotracheal tube should be changed.

18.1  Case 1

507

On the evening of the next day, the low pressure and low minute ventilation alarms began to sound. Also, the inspired tidal volume exceeded the expired tidal volume. The potential problem could be any of the following except: (a) ET obstruction (b) ET cuff leak or rupture (c) Air leak from temperature monitor port (d) Leak from exhalation valve The correct answer is (a). Note that this is an “except” question. Endotracheal tube obstruction leads to the activation of the high airway pressure alarm and not the low airway pressure alarm. The other three mentioned above can all lead to a fall in airway pressures. Auscultation over the trachea revealed harsh breath sounds over the entire duration of the inspiratory phase. What should be done? (a) Increase the minute volume (b) Check and inflate pilot bulb to the required pressure (c) Increase flow rate (d) All of the above The correct answer is (b). The presence of a hiss over the trachea during the ventilator delivered breath argues in favor of a cuff leak. The cuff should be reinflated by inflating the pilot bulb. If the pilot bulb fails to fill despite inflation with repeated boluses of air, a cuff rupture is likely and the ET should be replaced. The ET was replaced, but with a much smaller sized endotracheal tube owing to some difficulty in reintubation. The inflation of the pilot bulb resulted in the disappearance of the inspiratory sound, but the pilot bulb pressure required to achieve this seal was 36 cm H2O. The problem now was: (a) Right main-stem intubation (b) Size of ET too small for the patient’s airway

508

Chapter 18.  Case Studies

(c) Defective pilot bulb (d) All of the above The correct answer is (b). Much more air needs to be introduced into the pilot bulb of an ETT which has a diameter that is small relative to the trachea, in order to ensure an effective seal.

18.2 Case 2 Mrs. B, aged 32 years was brought to the ICU after falling off the pillion of a motor-scooter half an hour ago. She had become unconscious after the fall and had not gained consciousness since. On arrival at the ICU, she had a BP of 140/90 and a heart rate of 108 beats/min. She had a single bruise on her occiput, and there were multiple lacerations on her chest and arms. Her lungs were clear to auscultation. Withdrawal response to painful stimuli was present and the pupils were equal and responsive. A CT scan of the brain showed an intracranial bleed. Mrs. B’s chest film showed fractures of three ribs on her right side, as well as a small right sided pneumothorax. An intracranial probe revealed that the intracranial pressure was raised. The indications for intubation in the case of Mrs. B would be: (a) To protect the airway (b) To ventilate with an intent to produce hypocapnia (c) Both the above (d) None of the above The correct answer is (c). In cases of neurological injury, mechanical ventilation with deliberate hyperventilation with the intent of producing hypocapnia is not resorted to unless the intracranial pressure is high. In cases of raised intracranial pressure, hypocapnia has been gainfully employed to quickly reduce the intracranial tension. In a patient with poor

18.2  Case 2

509

airway defense mechanisms, intubation can protect the airway and prevent aspiration. Since even a small pneumothorax has the potential to transform itself into a tension pneumothorax on the ventilator, a chest drain was inserted into the right pleural cavity. A while later, expiratory tidal volumes fell well below the inspiratory tidal volumes. A circuit check revealed no obvious leak in the system. The most likely cause of this phenomenon could be: (a) Herniation of the endotracheal tube cuff (b) Alveolar instability (c) A large air leak through a bronchopleural fistula (d) Any of the above The correct answer is (c). Sustained bubbling from the underwater seal of the chest drain throughout the respiratory cycle was seen, which was surprising since the pneumothorax had been quite small. A large bronchopleural fistula was diagnosed. Appropriate action should now include: (a) Measures to reduce alveolar distension (b) High frequency jet ventilation (c) Sealing of the local bronchopulmonary segment by gelfoam or fibrin (d) Laser coagulation of the leak In theory, all these options have been attempted to treat bronchopleural fistulae. In practice, it is seldom necessary to do anything other than to limit the alveolar over-distension that often engenders them. Bronchopleural fistulae generally resolve parri-passu with improvement in the underlying lung pathology and in instances such as above where the minute ventilation is not compromised, a conservative line of management is often sufficient. Therefore, (a) is the correct answer.

510

Chapter 18.  Case Studies

18.3 Case 3 Mr. C, 56-years old, having chronic bronchitis, was admitted with an acute exacerbation of his condition, and in due course, had to be intubated and ventilated for hypercapnic respiratory failure which had progressed despite optional medical therapy. On day 2 of ventilatio on the assist-control mode, Mr C’s already elevated peak airway pressure suddenly rose further. Mr. C became tachypneic, began overbreathing the set respiratory rate, and looked distressed. A suction catheter passed down the endotracheal tube encountered no perceptible resistance and airway secretions were not much in evidence. Breath entry appeared markedly diminished on the left side. Diagnostic possibilities would include: (a) Collapse of the left lung due to secretions (b) Collapse of the left lung due to endotracheal tube migration into the right main-stem bronchus (c) Left sided pneumothorax (d) Any of the above The correct answer is (d). Lack of aspirable secretions from the endotracheal tube cannot rule out mucus plugging more distally, and endotracheal tube migration into the right main bronchus can obstruct the orifice of the left main bronchus causing absorbtive atelectasis. Barotrauma in a patient on mechanical ventilation is also a possibility at any time and a high degree of suspicion for the same must be maintained. An urgent bedside chest film revealed a collapsed left lung. The tip of the endotracheal tube was visualized well proximal to the carina and distal migration of the endotracheal tube was so ruled out. SpO2 at this stage was 92% on a FIO2 of 0.6. Appropriate therapy at this stage would be: (a) Chest physiotherapy (b) Urgent bronchoscopic toilet

18.4  Case 4

511

(c) Chest physiotherapy followed by bronchoscopic toilet, should the lung not expand with the former (d) Deep suctioning The correct answer is (c). Deep suctioning is not an option. In any case, a suction catheter would be liable to follow the path of least resistance and pass into the right main bronchus which is more aligned with the trachea and not the left main bronchus which is likely to be obstructed with a mucus plug or a blood clot. Chest therapy is a reasonable initial option since a saturation of 92% is for the time being satisfactory, but this should be followed up with a bronchoscopic toilet, should the lung not expand with chest physiotherapy alone.

18.4 Case 4 Mrs. D, aged 72 years was intubated and ventilated for type 1 respiratory failure secondary to cardiogenic pulmonary edema. Having improved, she was extubated after 48  h. Immediately post extubation, Mrs. D developed stridor and considerable respiratory difficulty. Diagnostic possibilities include: (a) Laryngospasm (b) Laryngeal edema (c) Either a or b (d) Tracheal stenosis. The correct answer is (c). Since the stridor occurred within a short time of extubation, laryngeal edema (as a result of injury to the larynx during intubation) could be proposed as a likely cause of stridor. Laryngospasm is also possible; tracheal stenosis usually takes much longer to manifest. Inspection of the larynx revealed laryngeal edema, but absence of other significant injury. At this stage, the initial treatment could include: (a) Tracheostomy (b) Corticosteroids

512

Chapter 18.  Case Studies

(c) Lidocaine (d) Racemic epinephrine, followed by reintubation if required, with a smaller sized endotracheal tube The correct answer is (d). Tracheostomy is not an option at this stage without allowing a little time for resolution of the laryngeal edema. Corticosteroids have not been convincingly demonstrated to be beneficial and although IV or tropical lidocaine has been used in cases of laryngospasm with some benefit, its role in laryngeal edema is unknown. Mrs. D required reintubation, and a smaller sized endotracheal tube was used to negotiate the swollen larynx. She was extubated after another 48  h and made an uneventful recovery. What parameter or test could have predicted the presence of laryngeal edema before extubation? (a) SpO2 (b) Tidal volume (c) Spontaneous respiratory effort (d) Cuff leak test The correct answer is (d). Deflation of the ET cuff by evacuating the pilot bulb should result in an inspiration sound on auscultation over the trachea, as air leaks between the ET tube and the airway. Lack of such a sound would imply a snug fit between the ET and the airway, and the cause of this could possibly be laryngeal edema.

18.5 Case 5 On day 7 of ventilation for a neuromuscular problem, Ms. E, aged 22 years, suddenly developed ventilatory distress. She was removed from the ventilator and bagged. A suction catheter passed down the ET encountered no resistance, and on auscultation, the breath entry was normal and equal. The therapist then hooked her back onto the ventilator, and noticed that although the oxygen saturation had fallen, the

18.6  Case 6

513

lung compliance and resistance were unchanged. The possible problem could be: (a) Pneumothorax (b) Pulmonary embolism (c) Lobar collapse (d) Endotracheal tube obstruction The correct answer is (b). Although it is possible for a small pneumothorax to go undetected on clinical examination, pneumothoraces in patients on mechanical ventilation often become tension pneumothoraces and become clinically obvious. Endotracheal obstruction is virtually ruled out by the unhindered passage of a suction catheter down the breathing tube, and the absence of clinical wheeze or a rise in airway pressure can reasonably exclude bronchospasm. Likewise, the absence of a fall in pulmonary compliance and/or a rise in airway pressures argues against major lobar atelectasis. Pulmonary embolism is the correct answer since it can produce hypoxemia without a significant change in either compliance or resistance. An unchanged chest film in this case strengthened the suspicion of pulmonary embolism which was later confirmed.

18.6 Case 6 Mr. F, a 42-year-old persistent asthmatic was intubated and ventilated during an attack of acute severe asthma which failed to respond to conventional medical therapy. He was a hypertensive, well controlled on medication and was in the habit of taking 5  mg diazepam nocte for the last 10 years. Postintubation ventilator settings were: assist-control mode with a backup of 14 breaths/min, at 500  mL/breath, FIO2 (which, a few minutes earlier had been 1.0) 0.6. After awakening from the sedation that he required at intubation, Mr. F became increasingly agitated and violent and had to be sedated and paralyzed. At this point, his pulse rate rose to 130

514

Chapter 18.  Case Studies

beats/min, respirations 36 breaths/min. SpO2 was 98%. Possible causes of hiss restlessness could be: (a) A blocked endotracheal tube (b) Steroid psychosis (c) Intrinsic PEEP (d) Any of the above The correct answer is (d). A blocked endotracheal tube is in many cases the cause of sudden restlessness and agitation in a ventilated patient. Tenacious secretions coat the luminal surface of the endotracheal tube and can progressively reduce the lumen of the endotracheal tube before a plug of inspissated mucus or clot of blood causes sudden total occlusion. The patient’s distress may not immediately be accompanied by a fall in SpO2, as a change in saturation usually takes time to register on the monitor. An acute rise in peak pressure is often the clue and resistance to the passage of a suction catheter down the endotracheal tube is virtually diagnostic of a tube block. Although ET blockage is unusual in a recently intubated patient, it should still be ruled out in such a situation. Sleep deprivation is also known to produce agitation and some patients on glucocorticoids develop steroid psychosis. The development of intrinsic-PEEP (autoPEEP) can lead to considerable patient-ventilator asynchrony and this can manifest in the patient fighting the ventilator. In this instance, the cause of the patient’s agitation proved to be a high intrinsic PEEP due to air-trapping and dynamic hyperinflation. Appropriate corrective measures would include all, except: (a) Bronchodilation (b) Decreasing I:E ratio (c) Increasing the duration of the inspiratory pause (d) Adding a small amount of external PEEP (about 50–75% of the measured intrinsic-PEEP) The correct answer is (c). Note that this is an “except” question. Decreasing inspiratory time leaves more time for

18.6  Case 6

515

expiration, and more complete lung emptying is possible, reducing dynamic hyperinflation, and thereby decreasing the auto-PEEP. The addition of a small amount of external PEEP reduces the gradient against which the patient must inspire, thus reducing the work of breathing. Increasing the pause time, however, (note that the inspiratory pause is considered part of the inspiratory time) will actually leave less time for expiration and will actually worsen dynamic hyperinflation. In Mr. F’s case, despite the application of optimal ventilator strategies, peak and plateau pressures remained at unacceptable levels; the setting of maximal airway pressure at 45 cm H2O resulted in termination of the inspiratory breath when 280 mL or so of the tidal volume had been delivered at each breath, and the hypoventilation resulted in a rise in the PaCO2 to 72 mmHg. Appropriate strategies at this stage comprise all of the following except: (a) Ventilation with helium–oxygen gas mixture (b) Pressure control ventilation (c) Permissive hypercapnia (d) Raising the upper airway pressure limit The correct answer is (d). Note that this is an “except” question. Since the set tidal volume is 500 mL and the breath is anyway being pressure limited to 280 mL, reduction in tidal volumes would serve no purpose. Since the low tidal volumes are now resulting in reduced minute ventilation, one way to allow delivery of the targeted tidal volumes would be to raise the upper airway pressure limit. This would, however, allow the upper airway pressure to rise and thereby unacceptably increase the risk of barotrauma. The other strategies are all acceptable in this situation, though the usage of heliox is limited to a few centers and not yet in the realms of conventional ventilatory strategy.

516

Chapter 18.  Case Studies

18.7 Case 7 Mr. G, a 21-year-old man was admitted to the ICU, after a collision with a bus caused his vehicle to overturn. On examination, Mr. G was well oriented and alert but in considerable pain. His heart rate was 110 beats/min, respirations 22 breaths/min and BP 140/80. Breath entry was equal and satisfactory on both sides, but a flail segment of the sternum moved paradoxically with each breath. The chest film confirmed the presence of multiple rib fractures bilaterally and some subcutaneous emphysema, but no obvious pneumothorax. The ABG on FIO2 of 0.4 showed pH 7.36, PaCO2 38, and PaO2 120  mmHg. Appropriate action now would be: (a) Observation and adequate analgesia (b) External stabilization of the chest by splints (c) Internal stabilization of the chest by intubation ventilation and the application of PEEP (d) Closed chest drainage Mr. G’s blood gases do not show the presence of respiratory failure, which is the indication for mechanical ventilation in flail chest. A PaCO2 of 38  mmHg is reasonable: it certainly does not indicate hypoventilation. The respiratory rate is a bit high, which is understandable since the impaired mechanics of the chest wall do not allow complete excursions of the chest and Mr. G is fulfilling his minute volume requirements by raising his respiratory rate; the latter by itself, is not high enough to impose a significantly high work of breathing. In fact, noninvasive ventilation could be considered if the work of breathing were judged to be bordering on high. Subcutaneous emphysema does not always equate with pneumothorax, though the chest X-ray should be carefully scrutinized for a small inobvious pneumothorax; it is also mandatory in such cases to closely watch the patient for the subsequent development of a “late” pneumothorax. External splints are often not very effective. Analgesia plays an important role, facilitates bronchial toilet, and prevents chest infections. The correct answer therefore is (a).

18.8  Case 8

517

Later on, Mr. G became increasingly tachypneic and distressed, and it was discovered that he did eventually develop a pneumothorax on the right. Closed chest drainage was performed, but during the while, Mr. G’s saturations dropped and he had to be intubated and ventilated. A postintubation chest film showed that the pneumothorax had resolved, but the left lung had now collapsed. Diagnostic possibilities would include: (a) Obstruction of the left main bronchus by clot (b) Obstruction of the left main bronchus by a mucus plug (c) Fracture of the left main bronchus (d) Any of the above The correct answer is (d). All three are possible in this setting. A diagnostic bronchoscopy discovered a mucus plug, the origin of which was uncertain, and the procedure proved to be of therapeutic benefit as well. Mr. G thereafter made an uneventful recovery.

18.8 Case 8 Mr. H, a 60-year-old diabetic, hypertensive, and smoker of thirty cigarettes per day for the last 40 years, was admitted with chest discomfort and difficulty in breathing since the last hour. On arrival, Mr. H was orthopneic, diaphoretic, and restless. His BP was recorded as 160/100, his heart rate 110 beats/ min, and respirations 40 breaths/min. The JVP was raised, there was bilateral pedal edema, and profuse basal crepitations were heard in both lungs. A chest film demonstrated cardiomegaly and bilateral symmetrical pulmonary parenchymal shadowing characteristic of cardiogenic pulmonary edema. The ECG showed evidence of evolving myocardial infarction. PaO2 remained at 47 mmHg in spite of high flow oxygen, and there was lack of significant response to diuresis and other medication. Treatment options would include: (a) Further observation and more diuresis (b) NIPPV

518

Chapter 18.  Case Studies

(c) Intubation and ventilation (d) All of the above The correct answer is (c), since the patient already satisfies a number of criteria for intubation and ventilation viz., refractory hypoxemia, tachypnea, and increased work of breathing in the presence of cardiac ischemia, especially in the face of lack of response to medication. It is unlikely that NIPPV would be beneficial at this stage since it would likely be ineffective in supporting the patient at this stage. The following is accepted as a conventional ventilatory strategy in CCF: (a) High frequency jet ventilation (b) Nitric oxide (c) PEEP (d) Heliox The correct answer is (c). The benefits that occur due to the application of PEEP are due to its ability to reduce preload in a failing heart. Also, by decreasing transmural aortic pressure, PEEP improves cardiac output, and by increasing functional residual capacity, it improves oxygenation and compliance. Since the effect of PEEP in an individual patient is by and large unpredictable, the patient should be closely monitored.

18.9 Case 9 Mrs. J, a 28-year-old asthmatic was brought to the ICU in status asthmaticus. She was breathless and tachypneic, with a respiratory rate of 28 breaths/min. Auscultation revealed markedly diminished breath entry on both sides and her chest film showed hyperinflated lungs, but no obvious infiltrate. An ABG taken 2 h after the administration of IV steroid, O2, and continuous nebulization showed pH 7.20, PaCO2 44, and PaO2 75 mmHg on FIO2 0.5. What should be the further course of action?

18.9  Case 9

519

(a) Continue treatment with continuous nebulization (b) Noninvasive positive pressure ventilation (c) Addition of a diuretic (d) Consider intubation and ventilation The correct answer is (d). A patient in exacerbation of asthma should normally be hyperventilating and will therefore show a reduced PaCO2. A normal PaCO2 implies that the partial pressure of CO2 is beginning to trend upward due to respiratory muscle exhaustion. In Mrs. J’s case, the acidic pH testifies to this. At this juncture, it is very unlikely that NIPPV will provide enough support to the respiratory muscles to reverse the critical process. The choice in such a patient should be elective intubation and ventilation when it becomes obvious that the PaCO2 is trending upward and the patient is fatiguing. After the intubation and ventilation, Mrs. J was put on the assist-control mode. All of the following ventilator settings would be appropriate in her case except: (a) Large tidal volumes to wash out the accumulating PaCO2 (b) Titration down of the FIO2 to keep PaO2 above 60 mmHg (c) See the upper airway pressure alarm at 40 cm H20 (d) Low I:E ratio The correct answer is (a). Note that this is an “except” question. Large tidal volumes in the setting of airway obstruction have the potential to exacerbate dynamic hyperinflation which is already a problem in such circumstances. A FIO2 tailored to keep PaO2 above 60 mmHg is sufficient, for at this PaO2 the hemoglobin should be near-completely saturated with O2. A low I:E ratio helps in that, with the shorter inspiration, a longer expiratory time is available to the overdistended lung to empty. A peak pressure of 40 and above has been linked to an increased risk of barotrauma, though it now appears that plateau pressures of more than 35  cm H2O equally, if not more closely, correlate with pressure-induced pulmonary injury.

520

Chapter 18.  Case Studies

18.10 Case 10 Mr. K, a COPD patient was ventilated 2 days ago for type 2 respiratory failure. At this stage, his current ventilator settings were as follows: On the assist-control mode, the tidal volume was 550, the respiratory rate 15 breaths/min, FIO2 0.28, inspiratory trigger −2 cm H2O, and inspiratory flow rate 50 L/min. Although the blood gases and vitals were acceptable, Mr. K evinced a sense of dyspnea. Which of the following changes in the ventilator settings would be likely to help? (a) Increasing the tidal volume (b) Increasing the respiratory rate (c) Increasing the FIO2 (d) Increasing the inspiratory flow rate The correct answer is (d). Since the blood gases are acceptable the minute ventilation need not be changed. Similarly, if the PaO2 is >60 mmHg on the present FIO2 (0.28), no change in FIO2 is required either. The inspiratory trigger is fairly low, so this should not impose a significant inspiratory load on Mr. K’s respiratory muscles. It is important, however, to realize that many persons, particularly patients with normal respiratory drives, require high inspiratory flow rates to fulfill the demands of their respiratory centers. Increasing the inspiratory flows should help. Later, Mr. K was put on SIMV (set rate 10 breaths/min) with a pressure support of 5 cm. The respiratory care practitioner noticed that Mr. K’s spontaneous tidal volumes ranged between 130 and 210 mL. What could be done to increase the tidal volume of the spontaneous breaths: (a) Increase the number of SIMV breaths (b) Increase the level of pressure support (c) Increase the tidal volume (d) All of the above

18.10  Case 10

521

The correct answer is (b). Increasing the level of pressure support can help increase the tidal volume of spontaneous breaths. A few days later Mr. K seemed better; he was still being ventilated on the SIMV with PS mode, at a pressure support of 10 cm H2O. Mr. K was then breathing at the SIMV backup rate of 9 breaths/min and taking an additional six breaths of his own at a tidal volume of 350–400 mL, through an endotracheal tube of size 7.5. In the morning, the ABG read as follows: pH 7.38, PCO2 32 mmHg, and PaO2 154 mmHg on 0.5 FIO2. On switching to the pressure support mode (PSV of 10  mmHg), Mr. K’s spontaneous tidal volumes fell to 250 mL and spontaneous respiratory frequency rose to 35 breaths/min, accompanied by subjective and objective signs of distress.The respiratory care practitioner reverted to the previous mode. Which of the following could make Mr. K wean successfully? (a) Reduction in FIO2 (b) Change in the endotracheal tube to a larger size (c) Progressively raising the PCO2 to approximately 50 mmHg before beginning the weaning process (d) All of the above The correct answer is (d). Theoretically, a reduction in FIO2 to 0.28 or thereabouts would help boost the respiratory drive of the chronic lunger who is habitually accustomed to a high PaCO2 though admittedly there appears nothing grossly wrong with Mr. K’s respiratory drive at this moment. By the same token, the starting point for a spontaneous breathing trial for a COPD patient should ideally be at a PaCO2 level that is at the patient’s usual premorbid baseline, and so a PaCO2 buildup to about 50 mmHg or even higher is considered appropriate in such cases before commencing the weaning trial. When weaning becomes difficult, a change in the endotracheal tube to as large a size as possible would help in substantially reducing the airway resistance and considerably help in the weaning process.

522

Chapter 18.  Case Studies

18.11 Case 11 Mr. L, a 72-year-old COPD patient was brought to the EMD. During transport, Mr. L had been given oxygen supplementation by a partial rebreathing mask at 12 L/min. At reception, Mr. L was drowsy and was breathing at only 5–6 breaths/min. An ABG performed on arrival showed pH 7.19, PaO2 66 mmHg, PaCO2 92 mmHg. Treatment options would include: (a) Administering FIO2 at 0.28 by ventimask (b) Making the patient breathe room air (c) Increasing FIO2 by administering O2 through a nonrebreathing mask (d) Intubation and ventilation The correct answer is (a). It is conceivable that the high flow oxygen administered to the patient during transport has suppressed his respiratory drive, compounding the hypercapnic respiratory failure. Reducing the FIO2 to an acceptable level is logical, in that the patient would still be getting enough FIO2 to maintain a reasonable O2 saturation and the reduction in FIO2 would also allow Mr. L’s hypoxic respiratory drive to improve his ventilatory status. Completely stopping supplemental oxygen is not an option. The PAO2 (the partial pressure of oxygen in the alveolus) is determined by the following equation: PAO2 = [(Atm pressure − Partial pressure of water vapour) × FIO2] − [(PaCO2 / respiratory quotient)] At sea level with a respiratory quotient of 0.8, assuming that 12 (LPM) by partial rebreathing mask corresponds to an FIO2 of approximately 0.6, PAO2 = [(760 – 47) × 0.6] – [(92/0.8)] = 312.8 mm Hg If the supplemental O2 were to be suddenly stopped, the patient would be breathing room air only (FIO2 > 0.21) and with a PCO2 of 92,

18.12  Case 12

523

PAO2 = [(760 – 47) × 0.21] – [(92/8)] = 34 mm Hg A low partial pressure of oxygen in the alveolus would mean that the arterial PaO2 would be lower yet, and this could result in cerebral hypoxia. Supplemental oxygen should therefore never be completely removed. Many physicians may feel that it would be hasty to intubate and ventilate the patient in a situation such as this, and would rather give the patient a chance to recover with a trial of initial conservative therapy. The role of NIPPV in this setting is unclear. Certainly, in a drowsy patient NIPPV is relatively contraindicated, (and here, there may be differences of opinion), but can be tried as the sensorium begins to improve. Bronchodilators, corticosteroids, antibiotics and respiratory stimulants may be used as the situation demands, with recourse to mechanical ventilation being taken if the PaO2 is not sustainable at ³60 mmHg with conservative therapy, or if there is a progressive rise in PaCO2 with acidosis inspite of optimal treatment.

18.12 Case 12 Mrs. M aged 30 years, weighing 65 kg is being ventilated for severe ARDS. The mode of ventilation initially used is assistcontrol. Appropriate tidal volumes for this patient should be: (a) 400 mL (b) 600 mL (c) 800 mL (d) 1,000 mL The correct answer is (a). Patients with severe ARDS have “baby lungs.” This means that healthy alveoli comprise approximately a third of the lung volume; another third of the lung is represented by collapsed but recruitable alveoli, while the remaining third is composed of densely consolidated or collapsed alveoli. Ventilating such patients with large tidal volumes would cause overdistension of the healthy

524

Chapter 18.  Case Studies

and compliant alveoli, resulting in alveolar injury. Low tidal volumes coupled with enough PEEP to hold the unstable alveolar units open is the currently recommended strategy for ARDS. At approximately 6–7  mg/kg body weight, a 400 mL tidal volume would be appropriate for Mrs. M. Mrs. M’s set respiratory rate is 22 breaths/min, FIO2 is 0.9 and PEEP is 5 cm H2O. Her BP is 140/90 and her ABG is as follows: pH 7.39 PaCO2 43  mmHg and PaO2 49  mmHg. An appropriate intervention to increase Mrs. R’s oxygenation would be: (a) Increase the FIO2 (b) Increase the PEEP (c) Increase the respiratory frequency (d) Lower the tidal volume further The correct answer is (b). The FIO2 is already 0.9, which is too high to be safe and the physician should try to reduce the FIO2 rather than increase it, in order to decrease the chances of oxygen-induced lung injury. Increasing the PEEP would be a good idea, since the blood pressure at this stage is well preserved. During the incremental application of PEEP to 14 cm H2O, monitoring of which one parameter is particularly relevant? (a) Heart rate (b) Blood pressure (c) Urine output (d) EtCO2 The correct answer is (b). Increase in PEEP decreases the venous return to the thorax and can have an early impact upon the blood pressure which therefore should be closely monitored. Mrs. M tolerated a PEEP of 14 cm well, and her BP was steady at 130/80. Over the next few days, Mrs. M’s lung mechanics worsened further and she was switched to pressure control ventilation. Despite all attempts to decrease the FIO2 to 0.6, it was not possible to do so and Mrs. M’s PaO2 was 50 on a FIO2 of 1.0

18.12  Case 12

525

and PEEP 16 cm H2O. The next option to increase Mrs. M’s oxygenation should be: (a) Permissive hypercapnia (b) Inverse ratio ventilation (c) ECCO2R (d) ECMO The correct answer is (b). PC-IRV (pressure control-inverse ratio ventilation) or VC-IRV (volume control-inverse ratio ventilation) should be the next option. Patients find the inversion of the respiratory time extremely uncomfortable and regularly require deep sedation with or without paralyzing agents. Placing the patient in a prone position too may help. The benefits conferred by extracorporeal life support are slender, though recent reports show better outcomes with LFPPV-ECCO2R. With the application of PC-IRV with a PCV level of 40  cm H2O and an I:E ratio 1:1, a PEEP of 12 cm H2O and an FIO2 of 0.8, the PaO2 was a bit better at 58 mmHg, but the PaCO2 gradually climbed to 55 with a progressive fall in the tidal volume. Further attempts to increase the I:E ratio resulted in a fall in the patient’s blood pressure. An appropriate option at this point of time would be: (a) Permissive hypercapnia (b) Increasing I:E ratio to 2:1 (c) High frequency jet ventilation (d) Increasing the PEEP The correct answer is (a). The problem now seems to be that the increasing airway pressure is limiting the tidal volumes causing the PaCO2 to rise, and compromising the hemodynamics. In the end, it may be better to accept a moderate rise in CO2 since the PaO2 seems better than before.

Index

A A-a DO2, 152–156 ABGs. See Arterial blood gases Accelerating waveform, 123 ACMV. See Volume assist-control mode Acute hypoxemic respiratory failure (AHRF), 424–426 Acute respiratory distress syndrome (ARDS) case study, 523–524 characterization criteria, 248–249 historical aspects, 5–6 pathophysiology diffuse alveolar damage (DAD), 251 Starling equation, 250 primary and secondary, 249 ventilatory strategies airway pressures, 253–254 flow waveforms, 254–255 inspiratory time, 259 inverse ratio ventilation, 259 overdistension, 256–258 PEEP, 255–256 permissive hypercapnia, 261–263 prone ventilation, 263–266 recruitment maneuvers, 260–261 respiratory rate, 254 tidal volumes, 253 ventilation modes, 252–253 Acute respiratory failure (ARF). See Noninvasive ventilation (NIV) Adaptive support ventilation (ASV), 109–110 Adjunctive therapies, 479–504 Adenosine triphosphate (ATP), 171 Adsorptive altelectasis, 326–327 Aerosolized antibiotics, 370–371

527

528

Index

Aerosol therapy, 486 Airway humidification. See Humidification, airway Airway occlusion pressure, 399 Airway pressure release ventilation (APRV), 97–98 Airway pressures. See Pressure-time scalar Airway resistance (Raw) calculation, 37, 38 Poiseuille’s law, 35 pulmonary elastance (Pel), 35–36 Alarms apnea, 147 high expired minute volume, 143–144 low airway pressure limit, 146 low expired minute volume activating conditions, 142, 143 airway alarm activation, 141, 142 low oxygen concentration (FIO2), 146 oxygen concentration, 146 power failure, 147 two-minute button, 148 upper airway pressure limit, 144–146 upper oxygen concentration (FIO2), 147 Alveolar dead-space, 40 Alveolar ventilation, 42–43 Anatomical dead-space, 40 Antibiotics aerosolized, 370–371 cycling, 374–375 resistance ESBL, 366 MRSA, 368 multi-drug resistance (MDR), 367 principles, 365 selection, 368–370 topical, 362–363 Apnea alarm, 147 APRV. See Airway pressure release ventilation Arrhythmias, 244, 311–312 Arterial blood gases (ABGs), 159 Arterial oxygen tension A-a DO2, 152–156 oxygen cascade, 149–150 PaO2, 151 PaO2/FIO2 ratio, 151

PaO2/PAO2 ratio, 151–152 Arteriovenous oxygen difference, 172–173 Aspiration, 312 Assist control mode, 84 Asthma. See Obstructive lung disease ASV. See Adaptive support ventilation Asynchrony, patient-ventilator, 225 flow asynchrony, 227–238 triggering asynchrony, 226–227 ventilator support level and work of breathing, 223 Automode, 106–107 Auto-PEEP clinical signs, 233, 234 flow-volume loop, 235 pressure-volume loop, 235, 236 measurement, 236–238 B BAL. See Broncho-alveolar lavage technique Barotrauma biotrauma, 321–322 case study, 510 manifestations, 318, 319 pneumomediastinum, 319–320 volutrauma, 320–321 Baseline variables, 78–79 Bilevel positive airway pressure (Bi-PAP), 97, 98 Biofilms, 349–350 Biotrauma, 321–322 Body suit, 442–443 Bohr equation, 45, 46 Broncho-alveolar lavage (BAL) technique, 358–359 Bronchopleural fistula (BPF), 278–279 Bronchopulmonary dysplasia, 333 Bronchospasm, 312Bulbar muscles involvement, 285–286 C Capnography components, 174–175 CPR, 182 mainstream and sidestream sensors, 180–181 PetCO2 factors affecting, 177

Index

529

530

Index

health and disease, 176 vs. PaCO2, 179 sidestream analyzers, 178 Carboxyhemoglobin (CO), 168 Cardiogenic pulmonary edema, 517–518 Cardiopulmonary resuscitation (CPR), 182 Cerebral autoregulation, 245, 246 Cerebral vasoconstriction, 247 Cervical spine injury, 308–309 Circuit, pneumatic 73 Compliance, lung dynamic, 32–34 respiratory system, 31 static, 29–31 Compliance, rate, oxygenation, and pressure (CROP) index, 404 Continuous positive airway pressure (CPAP), 95–97, 415 Control panel, 72 Control variables, 74–75 Cooperativity, 158 COPD exacerbation. See also Obstructive lung disease case study, 505 hypercapnic respitatory failure, 426–427 Cough strength, assessment, 409–410 CO2 transport, 157 CPAP. See Continuous positive airway pressure CPR. See Cardiopulmonary resuscitation Critical delivery of oxygen (DO2 crit), 174 CROP. See Compliance, rate oxygenation, and pressure index Cuff leak, 315–317 Cuirass, 3–4, 443 Cycle variables flow-cycled breath, 77 pressure-cycled breath, 78 time-cycled breath, 78 volume-cycled breath, 76–77 respiratory cycle phases, 74 D Dead-space alveolar, 40 anatomical, 40 physiological, 40–46 Decelerating waveform, 123 Diffuse alveolar damage, 332–333

DO2 crit. See Critical delivery of oxygen Dual breath control, 102–103 adaptive support ventilation (ASV) , 109–110 automode, 106–107 interbreath control, 103 intrabreath control, 102–103 mandatory minute ventilation (MMV), 107–108 pressure regulated volume control (PRVC), 103–106 advantages, 105–10 algorithm, 104 disadvantages, 106 modes, 105 volume support (VS), 108–109Dynamic hyperinflation causation mechanism, 316 causes, 317 exhalation problems, 318 functional residual capacity (FRC), 315 treatment strategies, 317 E ECCO2R. See Extracorporeal CO2 removal ECLS. See Extracorporeal life support ECMO. See Extracorporeal membrane oxygenation Endotracheal tube (ET), 19–21 Endotracheal tube obstruction, 313 EPAP. See Expiratory positive airway pressure Epistaxis, 307–308 ESBL. See Extended-spectrum beta-lactamase Esophageal intubation, 309–310 Esophageal perforation, 310 ET. See Endotracheal tube Expiratory asynchrony, 231–238 Expiratory hold and expiratory retard, 79–80 Expiratory valve, 73 Expiratory positive airway pressure (EPAP), 416 Extended-spectrum beta-lactamase (ESBL), 366 Extracorporeal CO2 removal (ECCO2R), 487 Extracorporeal life support (ECLS), 486–488 Extracorporeal membrane oxygenation (ECMO), 486–487 Extubation airway evaluation, 410–411 strength of cough assessment, 409–410 technique, 410 Eye irritation, 431

Index

531

532

Index

F Flail chest, 289–290 Flow asynchrony constant flow-volume-targeted ventilation, 228 expiratory asynchrony auto-PEEP, 233–238 delayed termination, 231–232 premature termination, 232–234 flow and pressure volume loop, 230 flow-time scalar, 229 Flow profile, 122–123 Flow rate, 118–119 Flow-time scalar airflow obstruction, 199, 200 derived information, 197, 198 emphysema, 200 flow waveforms, 196 low compliance, 200, 201 Flow-volume loop airway secretions, 223, 225 auto PEEP, 219, 221 decreased compliance, 221 increased airway resistance, 217–220 pressure-control ventilation, 216, 218 pressure support ventilation (PSV), 216–217, 219 tubing compressibility, 222–224 volume loss, 221–222 volume-targeted ventilation, 215–217 Forced vital capacity (FVC), 12 FRC. See Functional residual capacity Functional residual capacity (FRC), 19 FVC. See Forced vital capacity G Gas exchange monitoring, 149–184 Gastric distension, 430 Gastrointestinal dysfunction, 63 H Haldane effect, 329, 330 Hand-washing, 360–361 Heated humidifiers (HHs), 453 Heat-moisture exchangers (HMEs), 455–456 Helium–oxygen mixtures, 493–494

Hemodynamic compromise, 431 Hemodynamic effects of mechanical ventilation, 55–58 Hemoglobin structure, 156, 158 oxygenated and non-oxygenated, 161 Hepatobiliary dysfunction, 62–63 HFJV. See High-frequency jet ventilation HFOV. See High-frequency oscillatory ventilation HFPPV. See High-frequency positive pressure ventilation HFPV. See High-frequency percussive ventilation High expired minute volume alarm, 143–144 High-frequency jet ventilation (HFJV), 482–484 High-frequency oscillatory ventilation (HFOV), 484–485 High-frequency percussive ventilation (HFPV), 485–486 High-frequency positive pressure ventilation (HFPPV), 482 Historical aspects of mechanical ventilation, 1–6 HMEs. See Heat-moisture exchangers Humidification, airway dry air, 451 overcondensation, 452 overheated air, 451–452 isothermic saturation boundary, 449–450 nasal mucosa role, 449 NIV, 456 water losses, 450 Hypercapnic respiratory failure case study, 510–511 COPD exacerbation, 426–427 decompensated obstructive sleep apnea, 427 Hyperoxic hypercarbia, 329–332Hypocapnia, 508 Hypovolemic shock, 244–245 Hypoxemia diffusion defect, 54–55 hypoventilation, 46–49 right to left shunt, 52–53 V/Q mismatch, 50–51 Hypoxemic respiratory failure, 424–426 Hypoxia, tissue, 172 I Indications for mechanical ventilation hypoventilation, 10–11 hypoxia, 9–10 increased work, breathing, 11

Index

533

534

Index

Inspiratory hold, 79 Inspiratory positive airway pressure (IPAP), 416 Interbreath control, 103 Intrabreath control, 102–103 Intracranial pressure, 246 Intermittent mandatory ventilation (IMV). See Synchronized intermittent mandatory ventilation Intrathoracic pressure (ITP), 21, 23 Intubation esophageal, 309–310 right main bronchial, 310–311 endotracheal, 305, 322, 425 Inverse ratio ventilation (IRV), 133, 479–480 IPAP. See Inspiratory positive airway pressure Iron lung. See Tank ventilator Isothermic saturation boundary, 449–450 ITP. See Intrathoracic pressure J Jacket ventilator, 442–443 Jet nebulizers, 466–468 L LaPlace’s law, 492 Laryngeal trauma, 306 Limit variable, 76 Liquid ventilation partial, 496–497 perfluorocarbons, 495 total liquid ventilation (TLV), 496 Low airway pressure limit alarm, 146 Lower inflection point (Pflex), 128 Low expired minute volume alarm activating conditions, 142, 143 airway alarm activation, 141, 142 Low oxygen concentration (FIO2) alarm, 146 pressure-volume loop airway pressures, 204–206 compliance, 208–212 elastic work and resistive work, 211–214 machine breath, 204, 213 PEEP, 206, 208 spontaneous breath, 203–204 transalveolar pressure gradient, 205–207



Index triggering, 206, 207 volume loss, 213–214 scalars flow-time, 196–201 pressure-time, 190–196 volume-time, 200–202 ventilator waveforms, 189

M Mandatory minute ventilation (MMV), 107–108 Mean inspiratory flow (Vt/Ti), 399 Metered-dose inhaler (MDI), 472 Methicillin-resistant Staphylococcus aureus (MRSA), 347, 368 Methicillin-sensitive Staphylococcus aureus (MSSA), 346 Microaspiration, 349 Migration, upwards, of endotracheal tube, 314 MMV. See Mandatory minute ventilation Monotherapy vs. combination therapy, VAP, 373 MRSA. See Methicillin-resistant Staphylococcus aureus Myocardial ischemia arrhythmias, 244 breathing effect and myocardial perfusion, 241, 242 PEEP potential effects, 241, 243 tidal volume effect, 242, 243 N NAVA. See Neurally adjusted ventilatory assist Nebulization, 469–471 Nebulizers jet nebulizers aerosolization, 466 factors affecting, 467 instrumentation, 464–465 technique, 467–468 ultrasonic, 468–469 vibrating mesh, 469 Nebulization, 469–471 Negative pressure ventilation (NPV), 441–445 body suit, 442–443 cuirass, 443 drawbacks, 445 modes, 444 principle, 441 tank ventilator, 442

535

536

Index

Neurally adjusted ventilatory assist (NAVA), 497 Neurological injury, 247 case study, 508–509 cerebral autoregulation, 245, 246 cerebral vasoconstriction, 247 intracranial pressure, 246 Neurologic complications, 312 Neuromuscular disease, 512–513 bulbar muscles involvement, 285–286 spinal injury, 280, 281 NIPPV. See Noninvasive positive pressure ventilation Nitric oxide (NO), 488–491 NIV. See Noninvasive ventilation Nonhomogenous lung disease, 288–289 Noninvasive positive pressure ventilation (NIPPV/NIV), 275, 409 Noninvasive ventilation (NIV) advantages, 416 air leaks, 422–424 bronchoscopy, 428 contraindications, 418, 432 CPAP, 415 devices, 421 extubation failure, 428 humidification, 422 hypercapnic respitatory failure, 426–427 hypoxemic respiratory failure, 424–426 indications, 427–432 initiation steps, 428–429 interfaces, 418–420 modes, 420–421 monitoring, 432 outcomes, 432–433 weaning, 427 NPV. See Negative pressure ventilation O Obstructive lung disease asthma and COPD, general treatment principles, 276 bronchopleural fistula, 278–279 ET size, 276–277 external PEEP, 275 general anesthesia, 277 NIV, 275 PaCO2, 268–269

permissive hypercapnia, 277 pressure support mode, 271–272 respiratory rate, 273 tidal volume, 272–273 trigger sensitivity, 275 ventilator settings, 272–275 Open-loop and closed-loop systems, 72 Open lung concept, 135 Optical plethysmography, 160–161 Oxygen cascade, 149–150 Oxygen concentration alarm, 146 Oxygen extraction ratio, 173 Oxygen-related lung injury, 326–333 Oxyhemoglobin dissociation curve, 163, 164 Otalgia, 430 P PaO2, 151 PaO2/FIO2 ratio, 151 PaO2/PAO2 ratio, 151–152 Patient-ventilator asynchrony flow asynchrony, 228 expiratory asynchrony, 231–238 flow and pressure volume loop, 230 flow-time scalar, 229 triggering asynchrony ineffective triggering, 227 response time, 226–227 trigger type, 227 ventilator support level and work of breathing, 223 PAV. See Proportional assist ventilation PCV. See Pressure-controlled ventilation Peak airway pressures, 190 PEEP. See Positive end expiratory pressure Phase variables, 75 Pharmacokinetics, VAP, 368–371 Pharyngeal trauma, 306 Plateau pressures, 190–191 Pneumatic nebulizers. See Jet nebulizers Pneumomediastinum, 319–320 Pneumothorax, 516–517 Poiseuille’s law, 20–21 Poncho-wrap, 442–443 Positive end expiratory pressure (PEEP)

Index

537

538

Index

advantages, 130–131 auto-PEEP overcoming, 126–127 barotrauma and lung injury protection, 125–126 case study, 524–525 disadvantages, 131 flow waveforms, 132–133 indications and forms, 127 inspiratory time, 133 prone ventilation, 134 titration, 128–130 PPB. See Positive pressure breathing Pressure, airway distending pressure, 25 intrapleural pressure, 21–26 intrathoracic pressure (ITP), 21, 23 mean airway pressure, 191–193 pause airway pressure, 320 peak aiway pressure, 190–191 pressure gradients, 24 transalveolar pressure, 205–207 transpulmonary pressure, 22 Pressure-controlled ventilation (PCV) advantages and disadvantages, 99 PAV, 101–102 volume control and pressure control modes, 100 Pressure regulated volume control (PRVC) advantages, 105–106 algorithm, 104 disadvantages, 106 modes, 105 Pressure support ventilation (PSV) advantages, 91, 93 airway pressures, 89–90 disadvantages, 92, 93, 408–409 patient-ventilator asynchrony, 93–94 weaning algorithm, 407–408 Pressure-time product (PTI), 404 Pressure-time scalar mean airway pressures, 191–193 peak airway pressures, 190 plateau pressures, 190–191 Pressure-volume loop airway pressures, 204–206 compliance



Index

low, 208–210 high, 211, 212 overdistension, lung, 211 elastic work and resistive work, 211–213 Pressurized metered-dose inhalers. See Metered-dose inhaler (MDI) Prone ventilation, 134, 480 nonconventional mode, 479–497 proning test, 264–265 Proportional assist ventilation (PAV), 101–102 PRVC. See Pressure regulated volume control PSV. See Pressure support ventilation Pulmo-wrap, 442–443 Pulse oximetry absorption spectra, 168 carboxyhemoglobin (CO), 168 cooperativity, 158 disadvantages, 163 error sources, 167 hemoglobin structure, 156, 158 leftward shift and rightward shift, 165 light absorbance, 162 limitations, 161 optical plethysmography, 160–161 oxygenated and non-oxygenated hemoglobin, 161 spectrophotometry principle, 160 R Rapid shallow breathing index (RSBI), 402 Ratio of inspiration to expiration (I:E ratio) physiological effects, 120 prolonging inspiratory time, 121 Raw. See Resistance, airway (Raw) Recruitment maneuvers, 260–261 Renal effects, mechanical ventilation, 60–62 Resistance, airway (Raw) calculation, 37, 38 Poiseuille’s law, 35 pulmonary elastance (Pel), 35–36 tracheobronchial tree, 37 Reverse ramp pattern. See Decelerating waveform Reynold number, 493–494 Right main bronchial intubation, 310–311 RSBI. See Rapid shallow breathing index

539

540

Index

S Scalars flow-time, 196–201 pressure-time, 190–196 volume-time, 200–202 Self-extubation, 314–315 Simplified weaning index (SWI), 403 SIMV. See Synchronized intermittent mandatory ventilation Sine waveform, 123 Sinusitis, 322–323 complications, 354 nosocomial sinusitis treatment, 364 occurrence, 352 pathologic mechanisms, 353 Skin ulceration, 430 Spacers, 472–473 Spectrophotometry principle, 160 Spinal injury, 280, 281 Square waveform, 122 Spectrophotometry principle, 160Starling equation, 250 Stress ulcer prophylaxis, 362 Surfactant therapy, 491–493 SWI. See Simplified weaning index Synchronized intermittent mandatory ventilation (SIMV) advantages, 87–88 assist control mode vs. SIMV mode, 86–87 case study, 521 disadvantages, 88 weaning, 406–407 T Tank ventilator, 442 Time constants, mechanical ventilation, 38–39 Tissue oxygenation monitoring, 171–174 TLV. See Total liquid ventilation Tooth trauma, 308 Total liquid ventilation (TLV), 496 T-piece weaning, 4045–406Tracheal/bronchial rupture, 307 Tracheobronchitis, 328 Tracheocutaneous fistula, 326–327 Tracheoesophageal fistula air leakage, 323–324 cuff pressures, 325 formal repair, 326

Tracheoinnominate artery fistula, 325–326 Transcutaneous blood gas monitoring, 169–171 Transpulmonary pressure (PTA), 22 Triggering asynchrony, 226–227 Trigger variable, 75 Two-minute button alarm, 148 Type 1 respiratory failure, 511–512 Type-2 respiratory failure, 505–508, 520–521 U Ultrasonic nebulizers, 468–469 Upper airway pressure limit alarm, 144–146 Upper oxygen concentration (FIO2) alarm, 147 V VAP. See Ventilator-associated pneumonia Veno-arterial (VA) bypass-ECMO, 486–487 Veno-venous (VV) bypass-ECMO, 487 Ventilation and perfusion, 27 Ventilator alarms. See Alarms Ventilator-associated lung injury (VALI), 318–322 Ventilator-associated pneumonia (VAP) diagnosis differential, 355–356 sample interpretation, 358–360 sampling methods, 357–358 endotracheal tube/ventilator circuit, 363–364 incidence, 345 microbiology, 345–347 patient positioning, 354–355 prevention, 361–363 risk factors, 347–354 treatment, 365–374 antibiotic resistance, 365–368 duration, 371–373 lack of response, 373–374 pharmacokinetics, 368–371 Ventilator-induced lung injury (VILI), 318–322 Ventilator settings ET size, 276–277 flow profile, 122–123 flow rate, 118–119 fraction of inspired oxygen (FIO2), 131 I:E ratio, 120–121

Index

541

542

Index

open lung concept, 135 PEEP, 125–131 advantages, 130–131 auto-PEEP overcome, 126–127 barotrauma and lung injury protection, 125–126 disadvantages, 131 flow waveforms, 132–133 indications and forms, 127 inspiratory time, 133 IRV, 133 oxygenation, 124–125 oxygen carrying capacity, 134 oxygen consumption, 134 prone ventilation, 134 titration, 127–130 Vibrating mesh nebulizers (VMNs), 469 VMN. See Vibrating mesh nebulizers Volume assist-control mode (ACMV) advantages, 81–83 disadvantages, 83 trigger sensitivity, 82 Volume support (VS), 108–109 Volume-targeted modes. See Volume assist-control mode (ACMV) Volume-time scalar, 200–202 Volutrauma, 320–321 W Waveforms, ventilator, 189 Weaning central respiratory drive assessment, 399 factors affecting, 392 indices, 393–394 integrative indices CROP, 404 PTI, 404 RSBI f/Vt ratio, 402 SWI, 403 methods extubation, 409–411 NIPPV, 409 PSV, 407–409 synchronized IMV, 406–407 T-piece breathing, 405–406 oxygenation adequacy assessment



Index A-a DO2 gradient, 396 oxyhemoglobin dissociation curve, 395 PaO2/FIO2 ratio, 395–396 PaO2/PAO2 ratio, 396 respiratory muscle strength assessment minute ventilation, 398 PImax, 396–397 respiratory rate, 398 vital capacity, 397–398 respiratory system compliance, 401 sleep deprivation effects, 393 work of breathing, 400–401

543